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 78 毫秒
1.
2.
Relatively little is known about the small subset of peroxisomal proteins with predicted protease activity. Here, we report that the peroxisomal LON2 (At5g47040) protease facilitates matrix protein import into Arabidopsis (Arabidopsis thaliana) peroxisomes. We identified T-DNA insertion alleles disrupted in five of the nine confirmed or predicted peroxisomal proteases and found only two—lon2 and deg15, a mutant defective in the previously described PTS2-processing protease (DEG15/At1g28320)—with phenotypes suggestive of peroxisome metabolism defects. Both lon2 and deg15 mutants were mildly resistant to the inhibitory effects of indole-3-butyric acid (IBA) on root elongation, but only lon2 mutants were resistant to the stimulatory effects of IBA on lateral root production or displayed Suc dependence during seedling growth. lon2 mutants displayed defects in removing the type 2 peroxisome targeting signal (PTS2) from peroxisomal malate dehydrogenase and reduced accumulation of 3-ketoacyl-CoA thiolase, another PTS2-containing protein; both defects were not apparent upon germination but appeared in 5- to 8-d-old seedlings. In lon2 cotyledon cells, matrix proteins were localized to peroxisomes in 4-d-old seedlings but mislocalized to the cytosol in 8-d-old seedlings. Moreover, a PTS2-GFP reporter sorted to peroxisomes in lon2 root tip cells but was largely cytosolic in more mature root cells. Our results indicate that LON2 is needed for sustained matrix protein import into peroxisomes. The delayed onset of matrix protein sorting defects may account for the relatively weak Suc dependence following germination, moderate IBA-resistant primary root elongation, and severe defects in IBA-induced lateral root formation observed in lon2 mutants.Peroxisomes are single-membrane-bound organelles found in most eukaryotes. Peroxin (PEX) proteins are necessary for various aspects of peroxisome biogenesis, including matrix protein import (for review, see Distel et al., 1996; Schrader and Fahimi, 2008). Most matrix proteins are imported into peroxisomes from the cytosol using one of two targeting signals, a C-terminal type 1 peroxisome-targeting signal (PTS1) or a cleavable N-terminal type 2 peroxisome-targeting signal (PTS2) (Reumann, 2004). PTS1- and PTS2-containing proteins are bound in the cytosol by soluble matrix protein receptors, escorted to the peroxisome membrane docking complex, and translocated into the peroxisome matrix (for review, see Platta and Erdmann, 2007). Once in the peroxisome, many matrix proteins participate in metabolic pathways, such as β-oxidation, hydrogen peroxide decomposition, and photorespiration (for review, see Gabaldon et al., 2006; Poirier et al., 2006).In addition to metabolic enzymes, several proteases are found in the peroxisome matrix. Only one protease, DEG15/Tysnd1, has a well-defined role in peroxisome biology. The rat Tysnd1 protease removes the targeting signal after PTS2-containing proteins enter the peroxisome and also processes certain PTS1-containing β-oxidation enzymes (Kurochkin et al., 2007). Similarly, the Arabidopsis (Arabidopsis thaliana) Tysnd1 homolog DEG15 (At1g28320) is a peroxisomal Ser protease that removes PTS2 targeting signals (Helm et al., 2007; Schuhmann et al., 2008).In contrast with DEG15, little is known about the other eight Arabidopsis proteins that are annotated as proteases in the AraPerox database of putative peroxisomal proteins (Reumann et al., 2004; Carter et al., 2004; Shimaoka et al., 2004), which, in combination with the minor PTS found in both of these predicted proteases (Reumann, 2004), suggests that these enzymes may not be peroxisomal. Along with DEG15, only two of the predicted peroxisomal proteases, an M16 metalloprotease (At2g41790), which we have named PXM16 for peroxisomal M16 protease, and a Lon-related protease (At5g47040/LON2; Ostersetzer et al., 2007), are found in the proteome of peroxisomes purified from Arabidopsis suspension cells (Eubel et al., 2008). DEG15 and LON2 also have been validated as peroxisomally targeted using GFP fusions (Ostersetzer et al., 2007; Schuhmann et al., 2008).

Table I.

Putative Arabidopsis proteases predicted or demonstrated to be peroxisomal
AGI IdentifierAliasProtein ClassT-DNA Insertion AllelesPTSLocalization EvidenceLocalization References
At1g28320DEG15PTS2-processing proteaseSALK_007184 (deg15-1)SKL>aGFPReumann et al., 2004; Helm et al., 2007; Eubel et al., 2008; Schuhmann et al., 2008)
Proteomics
Bioinformatics
At2g41790PXM16Peptidase M16 family proteinSALK_019128 (pxm16-1)PKL>bProteomicsReumann et al., 2004, 2009; Eubel et al., 2008)
SALK_023917 (pxm16-2)Bioinformatics
At5g47040LON2Lon protease homologSALK_128438 (lon2-1)SKL>aGFPReumann et al., 2004, 2009; Ostersetzer et al., 2007; Eubel et al., 2008)
SALK_043857 (lon2-2)Proteomics
Bioinformatics
At2g18080Ser-type peptidaseSALK_020628SSI>cBioinformatics(Reumann et al., 2004)
SALK_102239
At2g35615Aspartyl proteaseSALK_090795ANL>bBioinformatics(Reumann et al., 2004)
SALK_036333
At3g57810Ovarian tumor-like Cys proteaseSKL>aBioinformatics(Reumann et al., 2004)
At4g14570Acylaminoacyl-peptidase proteinCKL>bBioinformatics (peroxisome)(Reumann et al., 2004; Shimaoka et al., 2004)
Proteomics (vacuole)
At4g20310Peptidase M50 family proteinRMx5HLdBioinformatics(Reumann et al., 2004)
At4g36195Ser carboxypeptidase S28 familySSM>bBioinformatics (peroxisome)(Carter et al., 2004; Reumann et al., 2004)





Proteomics (vacuole)

Open in a separate windowaMajor PTS1 (Reumann, 2004).bMinor PTS1 (Reumann, 2004).cValidated PTS1 (Reumann et al., 2007).dMinor PTS2 (Reumann, 2004).PXM16 is the only one of the nine Arabidopsis M16 (pitrilysin family) metalloproteases (García-Lorenzo et al., 2006; Rawlings et al., 2008) containing a predicted PTS. M16 subfamilies B and C contain the plastid and mitochondrial processing peptidases (for review, see Schaller, 2004), whereas PXM16 belongs to M16 subfamily A, which includes insulin-degrading peptidases (Schaller, 2004). A tomato (Solanum lycopersicum) M16 subfamily A protease similar to insulin-degrading enzymes with a putative PTS1 was identified in a screen for proteases that cleave the wound response peptide hormone systemin (Strassner et al., 2002), but the role of Arabidopsis PXM16 is unknown.Arabidopsis LON2 is a typical Lon protease with three conserved domains: an N-terminal domain, a central ATPase domain in the AAA family, and a C-terminal protease domain with a Ser-Lys catalytic dyad (Fig. 1A; Lee and Suzuki, 2008). Lon proteases are found in prokaryotes and in some eukaryotic organelles (Fig. 1C) and participate in protein quality control by cleaving unfolded proteins and can regulate metabolism by controlling levels of enzymes from many pathways, including cell cycle, metabolism, and stress responses (for review, see Tsilibaris et al., 2006). Four Lon homologs are encoded in the Arabidopsis genome; isoforms have been identified in mitochondria, plastids, and peroxisomes (Ostersetzer et al., 2007; Eubel et al., 2008; Rawlings et al., 2008). Mitochondrial Lon protesases are found in a variety of eukaryotes (Fig. 1A) and function both as ATP-dependent proteases and as chaperones promoting protein complex assemblies (Lee and Suzuki, 2008). LON2 is the only Arabidopsis Lon isoform with a canonical C-terminal PTS1 (SKL-COOH; Ostersetzer et al., 2007) or found in the peroxisome proteome (Eubel et al., 2008; Reumann et al., 2009). Functional studies have been conducted with peroxisomal Lon isoforms found in the proteome of peroxisomes purified from rat hepatic cells (pLon; Kikuchi et al., 2004) and the methylotrophic yeast Hansenula polymorpha (Pln; Aksam et al., 2007). Rat pLon interacts with β-oxidation enzymes, and a cell line expressing a dominant negative pLon variant has decreased β-oxidation activity, displays defects in the activation processing of PTS1-containing acyl-CoA oxidase, and missorts catalase to the cytosol (Omi et al., 2008). H. polymorpha Pln is necessary for degradation of a misfolded, peroxisome-targeted version of dihydrofolate reductase and for degradation of in vitro-synthesized alcohol oxidase in peroxisomal matrix extracts, but does not contribute to degradation of peroxisomally targeted GFP (Aksam et al., 2007).Open in a separate windowFigure 1.Diagram of LON2 protein domains, gene models for LON2, PXM16, DEG15, PED1, PEX5, and PEX6, and phylogenetic relationships of LON family members. A, Organization of the 888-amino acid LON2 protein. Locations of the N-terminal domain conserved among Lon proteins, predicted ATP-binding Walker A and B domains (black circles), active site Ser (S) and Lys (K) residues (asterisks), and the C-terminal Ser-Lys-Leu (SKL) peroxisomal targeting signal (PTS1) are shown (Lee and Suzuki, 2008). B, Gene models for LON2, PXM16, DEG15, PED1, PEX5, and PEX6 and locations of T-DNA insertions (triangles) or missense alleles (arrows) used in this study. Exons are depicted by black boxes, introns by black lines, and untranslated regions by gray lines. C, Phylogenetic relationships among LON homologs. Sequences were aligned using MegAlign (DNAStar) and the ClustalW method. The PAUP 4.0b10 program (Swofford, 2001) was used to generate an unrooted phylogram from a trimmed alignment corresponding to Arabidopsis LON2 residues 400 to 888 (from the beginning of the ATPase domain to the end of the protein). The bootstrap method was performed for 500 replicates with distance as the optimality criterion. Bootstrap values are indicated at the nodes. Predicted peroxisomal proteins have C-terminal PTS1 signals in parentheses and are in light-gray ovals. Proteins in the darker gray oval have N-terminal extensions and include mitochondrial and chloroplastic proteins. Sequence identifiers are listed in Supplemental Table S2.In this work, we examined the roles of several putative peroxisomal proteases in Arabidopsis. We found that lon2 mutants displayed peroxisome-deficient phenotypes, including resistance to the protoauxin indole-3-butyric acid (IBA) and age-dependent defects in peroxisomal import of PTS1- and PTS2-targeted matrix proteins. Our results indicate that LON2 contributes to matrix protein import into Arabidopsis peroxisomes.  相似文献   

3.
4.
Osteoarthritis (OA) is a multidimensional health problem and a common chronic disease. It has a substantial impact on patient quality of life and is a common cause of pain and mobility issues in older adults. The functional limitations, lack of curative treatments, and cost to society all demonstrate the need for translational and clinical research. The use of OA models in mice is important for achieving a better understanding of the disease. Models with clinical relevance are needed to achieve 2 main goals: to assess the impact of the OA disease (pain and function) and to study the efficacy of potential treatments. However, few OA models include practical strategies for functional assessment of the mice. OA signs in mice incorporate complex interrelations between pain and dysfunction. The current review provides a comprehensive compilation of mouse models of OA and animal evaluations that include static and dynamic clinical assessment of the mice, merging evaluation of pain and function by using automatic and noninvasive techniques. These new techniques allow simultaneous recording of spontaneous activity from thousands of home cages and also monitor environment conditions. Technologies such as videography and computational approaches can also be used to improve pain assessment in rodents but these new tools must first be validated experimentally. An example of a new tool is the digital ventilated cage, which is an automated home-cage monitor that records spontaneous activity in the cages.

Osteoarthritis (OA) is a multidimensional health problem and a common chronic disease.36 Functional limitations, the absence of curative treatments, and the considerable cost to society result in a substantial impact on quality of life.76 Historically, OA has been described as whole joint and whole peri-articular diseases and as a systemic comorbidity.9,111 OA consists of a disruption of articular joint cartilage homeostasis leading to a catabolic pathway characterized by chondrocyte degeneration and destruction of the extracellular matrix (ECM). Low-grade chronic systemic inflammation is also actively involved in the process.42,92 In clinical practice, mechanical pain, often accompanied by a functional decline, is the main reason for consultations. Recommendations to patients provide guidance for OA management.22, 33,49,86 Evidence-based consensus has led to a variety of pharmacologic and nonpharmacologic modalities that are intended to guide health care providers in managing symptomatic patients. Animal-based research is of tremendous importance for the study of early diagnosis and treatment, which are crucial to prevent the disease progression and provide better care to patients.The purpose of animal-based OA research is 2-fold: to assess the impact of the OA disease (pain and function) and to study the efficacy of a potential treatment.18,67 OA model species include large animals such as the horse, goat, sheep, and dog, whose size and anatomy are expected to better reflect human joint conditions. However, small animals such as guinea pig, rabbit, mouse, and rat represent 77% of the species used.1,87 In recent years, mice have become the most commonly used model for studying OA. Mice have several advantageous characteristics: a short development and life span, easy and low-cost breeding and maintenance, easy handling, small joints that allow histologic analysis of the whole joint,32 and the availability of genetically modified lines.108 Standardized housing, genetically defined strains and SPF animals reduce the genetic and interindividual acquired variability. Mice are considered the best vertebrate model in terms of monitoring and controlling environmental conditions.7,14,15,87 Mouse skeletal maturation is reached at 10 wk, which theoretically constitutes the minimal age at which mice should be entered into an OA study.64,87,102 However, many studies violate this limit by testing mice at 8 wk of age.Available models for OA include the following (32,111 physical activity and exercise induced OA; noninvasive mechanical loading (repetitive mild loading and single-impact injury); and surgically induced (meniscectomy models or anterior cruciate ligament transection). The specific model used would be based on the goal of the study.7 For example, OA pathophysiology, OA progression, and OA therapies studies could use spontaneous, genetic, surgical, or noninvasive models. In addition, pain studies could use chemical models. Lastly, post-traumatic studies would use surgical or noninvasive models; the most frequently used method is currently destabilization of the medial meniscus,32 which involves transection of the medial meniscotibial ligament, thereby destabilizing the joint and causing instability-driven OA. An important caveat for mouse models is that the mouse and human knee differ in terms of joint size, joint biomechanics, and histologic characteristics (layers, cellularity),32,64 and joint differences could confound clinical translation.10 Table 1. Mouse models of osteoarthritis.
ModelsProsCons
SpontaneousWild type mice7,9,59,67,68,70,72,74,80,85,87,115,118,119,120- Model of aging phenotype
- The less invasive model
- Physiological relevance: mimics human pathogenesis
- No need for technical expertise
- No need for specific equipment
- Variability in incidence
- Large number of animals at baseline
- Long-term study: Time consuming (time of onset: 4 -15 mo)
- Expensive (husbandry)
Genetically modified mice2,7,25,40,50,52,67,72,79,80, 89,120- High incidence
- Earlier time of onset: 18 wk
- No need for specific equipment
- Combination with other models
- Time consuming for the strain development
- Expensive
Chemical- inducedMono-iodoacetate injection7,11,46,47,60,66,90,91,101,128- Model of pain-like phenotype
- To study mechanism of pain and antalgic drugs
- Short-term study: Rapid progression (2-7 wk)
- Reproducible
- Low cost
- Need for technical expertise
- Need for specific equipment
- Systemic injection is lethal
- Destructive effect: does not allow to study the early phase of pathogenesis
Papain injection66,67,120- Short-term study: rapid progression
- Low cost
- Need for technical expertise
- Need for specific equipment
- Does not mimic natural pathogenesis
Collagenase injection7,65,67,98- Short-term study: rapid progression (3 wk)
- Low cost
- Need for technical expertise
- Need for specific equipment
- Does not mimic natural pathogenesis
Non-invasiveHigh-fat diet (Alimentary induced obesity model)5,8,43,45,57,96,124Model of metabolic phenotype
No need for technical expertise
No need for specific equipment
Reproducible
Long-term study: Time consuming (8 wk–9 mo delay)
Expensive
Physical activity and exercise model45,73Model of post traumatic phenotype
No need for technical expertise
Long-term study: time consuming (18 mo delay)
Expensive
Disparity of results
Mechanical loading models Repetitive mild loading models Single-impact injury model7,16,23,24, 32,35,104,105,106Model of post traumatic phenotype
Allow to study OA development
Time of onset: 8-10 wk post injury
Noninvasive
Need for technical expertise
Need for specific equipment
Heterogeneity in protocol practices
Repetitive anesthesia required or ethical issues
SurgicalOvariectomy114Contested.
Meniscectomy model7,32,63,67,87 Model of post traumatic phenotype
High incidence
Short-term study: early time of onset (4 wk from surgery)
To study therapies
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Anterior cruciate ligament transection (ACLT)7,39,40,61,48,67,70,87,126Model of posttraumatic phenotype
High incidence
Short-term study: early time of onset (3-10 wk from surgery)
Reproducible
To study therapies
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Destabilization of medial meniscus (DMM)7,32,39,40Model of post traumatic phenotype
High incidence
Short-term study: early time of onset (4 wk from surgery)
To study therapies
The most frequently used method
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Open in a separate windowSince all animal models have strengths and weaknesses, it is often best to plan using a number of models and techniques together to combine the results.In humans, the lack of correlation between OA imaging assessment and clinical signs highlights the need to consider the functional data and the quality of life to personalize OA management. Clinical outcomes are needed to achieve 2 main goals: to assess the impact of the OA in terms of pain and function and to study the efficacy of treatments.65 Recent reviews offer few practical approaches to mouse functional assessment and novel approaches to OA models in mice.7,32,67,75,79,83,87, 100,120 This review will focus on static and dynamic clinical assessment of OA using automatic and noninvasive emerging techniques (
Test nameTechniquesKind of assessmentOutputSpecific equipment required
Static measurement
Von Frey filament testingCalibrated nylon filaments of various thickness (and applied force) are pressed against the skin of the plantar surface of the paw in ascending order of forceStimulus- evoked pain-like behavior
Mechanical stimuli - Tactile allodynia
The most commonly used test
Latency to paw withdrawal
and
Force exerted are recorded
Yes
Knee extension testApply a knee extension on both the intact and affected knee
or
Passive extension range of the operated knee joint under anesthesia
Stimulus-evoked pain-like behaviorNumber of vocalizations evoked in 5 extensionsNone
HotplateMouse placed on hotplate. A cutoff latency has been determined to avoid lesionsStimulus-evoked pain-like behavior
Heat stimuli- thermal sensitivity
Latency of paw withdrawalYes
Righting abilityMouse placed on its backNeuromuscular screeningLatency to regain its footingNone
Cotton swab testBringing a cotton swab into contact with eyelashes, pinna, and whiskersStimulus-evoked pain-like behavior
Neuromuscular screening
Withdrawal or twitching responseNone
Spontaneous activity
Spontaneous cage activityOne by one the cages must be laid out in a specific platformSpontaneous pain behavior
Nonstimulus evoked pain
Activity
Vibrations evoked by animal movementsYes
Open field analysisExperiment is performed in a clear chamber and mice can freely exploreSpontaneous pain behavior
Nonstimulus evoked pain
Locomotor analysis
Paw print assessment
Distance traveled, average walking speed, rest time, rearing
Yes
Gait analysisMouse is placed in a specific cage equipped with a fluorescent tube and a glass plate allowing an automated quantitative gait analysisNonstimulus evoked pain
Gait analysis
Indirect nociception
Intensity of the paw contact area, velocity, stride frequency, length, symmetry, step widthYes
Dynamic weight bearing systemMouse placed is a specific cage. This method is a computerized capacitance meter (similar to gait analysis)Nonstimulus evoked pain
Weight-bearing deficits
Indirect nociception
Body weight redistribution to a portion of the paw surfaceYes
Voluntary wheel runningMouse placed is a specific cage with free access to stainless steel activity wheels. The wheel is connected to a computer that automatically record dataNonstimulus evoked pain
Activity
Distance traveled in the wheelYes
Burrowing analysisMouse placed is a specific cage equipped with steel tubes (32 cm in length and 10 cm in diameter) and quartz sand in Plexiglas cages (600 · 340x200 mm)Nonstimulus evoked pain
Activity
Amount of sand burrowedYes
Digital video recordingsMouse placed is a specific cage according to the toolNonstimulus evoked pain
Or
Evoked pain
Scale of pain or specific outcomeYes
Digital ventilated cage systemNondisrupting capacitive-based technique: records spontaneous activity 24/7, during both light and dark phases directly from the home cage rackSpontaneous pain behavior
Nonstimulus evoked pain
Activity-behavior
Distance walked, average speed, occupation front, occupation rear, activation density.
Animal locomotion index, animal tracking distance, animal tracking speed, animal running wheel distance and speed or rotation
Yes
Challenged activity
Rotarod testGradual and continued acceleration of a rotating rod onto which mice are placedMotor coordination
Indirect nociception
Rotarod latency: riding time and speed with a maximum cut off.Yes
Hind limb and fore grip strengthMouse placed over a base plate in front of a connected grasping toolMuscle strength of limbsPeak force, time resistanceYes
Wire hang analysisSuspension of the mouse on the wire and start the timeMuscle strength of limbs: muscle function and coordinationLatency to fall grippingNone
(self -constructed)
Open in a separate windowPain cannot be directly measured in rodents, so methods have been developed to quantify “pain-like” behaviors. The clinical assessment of mice should be tested both before and after the intervention (induced-OA ± administration of treatment) to take into account the habituation and establish a baseline to compare against.  相似文献   

5.
Natural Infection of Burkholderia pseudomallei in an Imported Pigtail Macaque (Macaca nemestrina) and Management of the Exposed Colony     
Crystal H Johnson  Brianna L Skinner  Sharon M Dietz  David Blaney  Robyn M Engel  George W Lathrop  Alex R Hoffmaster  Jay E Gee  Mindy G Elrod  Nathaniel Powell  Henry Walke 《Comparative medicine》2013,63(6):528-535
Identification of the select agent Burkholderia pseudomallei in macaques imported into the United States is rare. A purpose-bred, 4.5-y-old pigtail macaque (Macaca nemestrina) imported from Southeast Asia was received from a commercial vendor at our facility in March 2012. After the initial acclimation period of 5 to 7 d, physical examination of the macaque revealed a subcutaneous abscess that surrounded the right stifle joint. The wound was treated and resolved over 3 mo. In August 2012, 2 mo after the stifle joint wound resolved, the macaque exhibited neurologic clinical signs. Postmortem microbiologic analysis revealed that the macaque was infected with B. pseudomallei. This case report describes the clinical evaluation of a B. pseudomallei-infected macaque, management and care of the potentially exposed colony of animals, and protocols established for the animal care staff that worked with the infected macaque and potentially exposed colony. This article also provides relevant information on addressing matters related to regulatory issues and risk management of potentially exposed animals and animal care staff.Abbreviations: CDC, Centers for Disease Control and Prevention; IHA, indirect hemagglutination assay; PEP, postexposure prophylacticBurkholderia pseudomallei, formerly known as Pseudomonas pseudomallei, is a gram-negative, aerobic, bipolar, motile, rod-shaped bacterium. B. pseudomallei infections (melioidosis) can be severe and even fatal in both humans and animals. This environmental saprophyte is endemic to Southeast Asia and northern Australia, but it has also been found in other tropical and subtropical areas of the world.7,22,32,42 The bacterium is usually found in soil and water in endemic areas and is transmitted to humans and animals primarily through percutaneous inoculation, ingestion, or inhalation of a contaminated source.8, 22,28,32,42 Human-to-human, animal-to-animal, and animal-to-human spread are rare.8,32 In December 2012, the National Select Agent Registry designated B. pseudomallei as a Tier 1 overlap select agent.39 Organisms classified as Tier 1 agents present the highest risk of deliberate misuse, with the most significant potential for mass casualties or devastating effects to the economy, critical infrastructure, or public confidence. Select agents with this status have the potential to pose a severe threat to human and animal health or safety or the ability to be used as a biologic weapon.39Melioidosis in humans can be challenging to diagnose and treat because the organism can remain latent for years and is resistant to many antibiotics.12,37,41 B. pseudomallei can survive in phagocytic cells, a phenomenon that may be associated with latent infections.19,38 The incubation period in naturally infected animals ranges from 1 d to many years, but symptoms typically appear 2 to 4 wk after exposure.13,17,35,38 Disease generally presents in 1 of 2 forms: localized infection or septicemia.22 Multiple methods are used to diagnose melioidosis, including immunofluorescence, serology, and PCR analysis, but isolation of the bacteria from blood, urine, sputum, throat swabs, abscesses, skin, or tissue lesions remains the ‘gold standard.’9,22,40,42 The prognosis varies based on presentation, time to diagnosis, initiation of appropriate antimicrobial treatment, and underlying comorbidities.7,28,42 Currently, there is no licensed vaccine to prevent melioidosis.There are several published reports of naturally occurring melioidosis in a variety of nonhuman primates (NHP; 2,10,13,17,25,30,31,35 The first reported case of melioidosis in monkeys was recorded in 1932, and the first published case in a macaque species was in 1966.30 In the United States, there have only been 7 documented cases of NHP with B. pseudomallei infection.2,13,17 All of these cases occurred prior to the classification of B. pseudomallei as a select agent. Clinical signs in NHP range from subclinical or subacute illness to acute septicemia, localized infection, and chronic infection. NHP with melioidosis can be asymptomatic or exhibit clinical signs such as anorexia, wasting, purulent drainage, subcutaneous abscesses, and other soft tissue lesions. Lymphadenitis, lameness, osteomyelitis, paralysis and other CNS signs have also been reported.2,7,10,22,28,32 In comparison, human''s clinical signs range from abscesses, skin ulceration, fever, headache, joint pain, and muscle tenderness to abdominal pain, anorexia, respiratory distress, seizures, and septicemia.7,9,21,22

Table 1.

Summary of reported cases of naturally occurring Burkholderia pseudomalleiinfections in nonhuman primates
CountryaImported fromDate reportedSpeciesReference
AustraliaBorneo1963Pongo sp.36
BruneiUnknown1982Orangutan (Pongo pygmaeus)33
France1976Hamlyn monkey (Cercopithecus hamlyni) Patas monkey (Erythrocebus patas)11
Great BritainPhilippines and Indonesia1992Cynomolgus monkey (Macaca fascicularis)10
38
MalaysiaUnknown1966Macaca spp.30
Unknown1968Spider monkey (Brachytelis arachnoides) Lar gibbon (Hylobates lar)20
Unknown1969Pig-tailed macaque (Macaca nemestrina)35
Unknown1984Banded leaf monkey (Presbytis melalophos)25
SingaporeUnknown1995Gorillas, gibbon, mandrill, chimpanzee43
ThailandUnknown2012Monkey19
United StatesThailand1970Stump-tailed macaque (Macaca arctoides)17
IndiaPig-tailed macaque (Macaca nemestrina)
AfricaRhesus macaque (Macaca mulatta) Chimpanzee (Pan troglodytes)
Unknown1971Chimpanzee (Pan troglodytes)3
Malaysia1981Pig-tailed macaque (Macaca nemestrina)2
Wild-caught, unknown1986Rhesus macaque (Macaca mulatta)13
Indonesia2013Pig-tailed macaque (Macaca nemestrina)Current article
Open in a separate windowaCountry reflects the location where the animal was housed at the time of diagosis.Here we describe a case of melioidosis diagnosed in a pigtail macaque (Macaca nemestrina) imported into the United States from Indonesia and the implications of the detection of a select agent identified in a laboratory research colony. We also discuss the management and care of the exposed colony, zoonotic concerns regarding the animal care staff that worked with the shipment of macaques, effects on research studies, and the procedures involved in reporting a select agent incident.  相似文献   

6.
Surprising Fitness Consequences of GC-Biased Gene Conversion: I. Mutation Load and Inbreeding Depression     
Sylvain Glémin 《Genetics》2010,185(3):939-959
GC-biased gene conversion (gBGC) is a recombination-associated process mimicking selection in favor of G and C alleles. It is increasingly recognized as a widespread force in shaping the genomic nucleotide landscape. In recombination hotspots, gBGC can lead to bursts of fixation of GC nucleotides and to accelerated nucleotide substitution rates. It was recently shown that these episodes of strong gBGC could give spurious signatures of adaptation and/or relaxed selection. There is also evidence that gBGC could drive the fixation of deleterious amino acid mutations in some primate genes. This raises the question of the potential fitness effects of gBGC. While gBGC has been metaphorically termed the “Achilles'' heel” of our genome, we do not know whether interference between gBGC and selection merely has practical consequences for the analysis of sequence data or whether it has broader fundamental implications for individuals and populations. I developed a population genetics model to predict the consequences of gBGC on the mutation load and inbreeding depression. I also used estimates available for humans to quantitatively evaluate the fitness impact of gBGC. Surprising features emerged from this model: (i) Contrary to classical mutation load models, gBGC generates a fixation load independent of population size and could contribute to a significant part of the load; (ii) gBGC can maintain recessive deleterious mutations for a long time at intermediate frequency, in a similar way to overdominance, and these mutations generate high inbreeding depression, even if they are slightly deleterious; (iii) since mating systems affect both the selection efficacy and gBGC intensity, gBGC challenges classical predictions concerning the interaction between mating systems and deleterious mutations, and gBGC could constitute an additional cost of outcrossing; and (iv) if mutations are biased toward A and T alleles, very low gBGC levels can reduce the load. A robust prediction is that the gBGC level minimizing the load depends only on the mutational bias and population size. These surprising results suggest that gBGC may have nonnegligible fitness consequences and could play a significant role in the evolution of genetic systems. They also shed light on the evolution of gBGC itself.GC-BIASED gene conversion (gBGC) is increasingly recognized as a widespread force in shaping genome evolution. In different species, gene conversion occurring during double-strand break recombination repair is thought to be biased toward G and C alleles. In heterozygotes, GC alleles undergo a kind of molecular meiotic drive that mimics selection (reviewed in Marais 2003). This process can rapidly increase the GC content, especially around recombination hotspots (Spencer et al. 2006), and, more broadly, can affect genome-wide nucleotide landscapes (Duret and Galtier 2009a). For instance, it is thought to play a role in shaping isochore structure evolution in mammals (Galtier et al. 2001; Meunier and Duret 2004; Duret et al. 2006) and birds (Webster et al. 2006). Direct experimental evidence of gBGC mainly comes from studies in yeast (Birdsell 2002; Mancera et al. 2008; but see Marsolier-Kergoat and Yeramian 2009) and humans (Brown and Jiricny 1987). However, associations between recombination and the nucleotide landscape and frequency spectra biased toward GC alleles provide indirect evidence in very diverse organisms (OrganismsDirect evidenceIndirect evidenceAchille''s heel evidenceReferencesYeastMeiotic segregation biasMancera et al. (2008)Mitotic and mitotic heteromismatch correction biasCorrelation between GC and recombinationBirdsell (2002)MammalsMitotic heteromismatch correction biasBrown and Jiricny (1987)Correlation between GC*/GC and recombinationDuret and Arndt (2008); Meunier and Duret (2004)Biased frequency spectrum toward GC allelesGaltier et al. (2001); Spencer et al. (2006)GC bias associated with high dN/dS near recombination hotspotBerglund et al. (2009; Galtier et al. (2009)BirdsCorrelation between GC and recombinationInternational Chicken Genome Sequencing Consortium (2004)TurtlesCorrelation between GC and chromosome sizeKuraku et al. (2006)DrosophilaCorrelation between GC and recombinationMarais et al. (2003)Biased frequency spectrum toward GC allelesGaltier et al. (2006)NematodesCorrelation between GC and recombinationMarais et al. (2001)GrassesCorrelation between GC and outcrossing/selfingGlémin et al. (2006)Correlation between GC* and recombination and outcrossing/selfingOutcrossing increases dN/dS for genes with high GC*Haudry et al. (2008)Green algaeCorrelation between GC and recombinationJancek et al. (2008)ParameciumCorrelation between GC and chromosome sizeDuret et al. (2008)Open in a separate windowThe impact of gBGC on noncoding sequences and synonymous sites has been studied in depth, especially because of confounding effects with selection on codon usage (Marais et al. 2001). More recently, Galtier and Duret (2007) pointed out that gBGC may also interfere with selection when affecting functional sequences. They argued that gBGC could leave spurious signatures of adaptive selection and proposed to extend the null hypothesis of molecular evolution. Indeed, gBGC can lead to a ratio of nonsynonymous (dN) over synonymous (dS) substitutions above one (Berglund et al. 2009; Galtier et al. 2009), i.e., a typical signature of positive selection (Nielsen 2005). This hypothesis has been widely debated for human-accelerated regions (HARs). These regions are extremely conserved across mammals but show evidence of accelerated evolution along the human lineage, which has been interpreted as evidence of positive selection (Pollard et al. 2006a,b; Prabhakar et al. 2006, 2008). On the contrary, other authors argued that patterns observed in HARs, such as the AT → GC substitution bias, the absence of a selective sweep signature, or the propensity to occur within or close to recombination hotspots, are more likely explained by gBGC rather than positive selection (Galtier and Duret 2007; Berglund et al. 2009; Duret and Galtier 2009b; but see also Pollard et al. 2006a who also suggested that gBGC might play a role in HARs evolution). It is thus crucial to take gBGC into account when interpreting genomic data.Moreover, Galtier and Duret (2007) initially suggested that gBGC hotspots could contribute to the fixation of slightly deleterious AT → GC mutations and could represent the Achilles'' heel of our genome. This hypothesis was reinforced later in primates, with evidence of gBGC-driven fixation of deleterious mutations in proteins (Galtier et al. 2009). A similar result was also found in some grass species, whose genomes are also supposed to be affected by gBGC (Glémin et al. 2006). Haudry et al. (2008) compared two outcrossing and two selfing grass species and showed that GC-biased genes exhibit higher dN/dS ratio in outcrossing than in selfing lineages. The reverse pattern would be expected under pure selective models because of the reduced selection efficacy in selfers (Charlesworth 1992; Glémin 2007). This pattern is in agreement with a genomic Achilles'' heel associated with outcrossing, while gBGC is inefficient in selfing species because they are mainly homozygous.Twenty years ago, Bengtsson (1990) already pointed out that biased conversion can generally affect the mutation load. The mutation load is the reduction in the mean fitness of a population due to mutation accumulation, which could lead to population extinction if it is too high (Lynch et al. 1995). At this time, Bengtsson concluded that “it is impossible to know if biased conversion plays a major role in determining the magnitude of the mutation load in organisms such as ourselves, but the possibility must be considered and further investigated (Bengtsson 1990, p. 186).” Now, one can propose gBGC could be such a widespread biased conversion process. It thus appears timely to thoroughly investigate the fitness consequences of gBGC through its potential effects on the dynamics of deleterious mutations. The fitness consequences of gBGC were also pointed out as a major future issue to be addressed by Duret and Galtier (2009a). In addition to the load, deleterious mutations have many other evolutionary consequences (for review see Charlesworth and Charlesworth 1998). They are thought to be the main determinant of inbreeding depression, i.e., the reduction in fitness of inbred individuals compared to outbred ones. They also play a key role in the evolution of genetic systems (sexual reproduction and recombination, inbreeding avoidance mechanisms, ploidy cycles), of senescence, or in the degeneration of nonrecombining regions, such as Y chromosomes. So far, we know little, if anything, about how gBGC might affect these processes.In his seminal work, Bengtsson (1990) did not address several important points. First, he did not include genetic drift in his model. Nearly neutral mutations, for which drift and selection are of similar intensities, are the most damaging ones because they can drift to fixation, unlike strongly deleterious mutations that are maintained at low frequency (Crow 1993; Lande 1994, 1998). While gBGC intensities are rather weak (Birdsell 2002; Spencer et al. 2006), they could markedly affect the fate of nearly neutral mutations (see also Galtier et al. 2009). Second, Bengtsson did not study the effect of gene conversion on inbreeding depression, while he showed that recessive mutations, mostly involved in inbreeding depression, are the most affected by gene conversion. Third, he did not envisage systematic GC bias with its opposite effects on A/T and G/C deleterious alleles. Fourth, while he noted that selfing affects both the efficacy of selection and that of conversion, he did not fully investigate the effect of mating systems. On one hand, selfing is efficient in purging strongly deleterious mutations causing inbreeding depression. However, since selfing is expected to increase drift, weakly deleterious mutations can fix in selfing species, contributing to the so-called “drift load” (Charlesworth 1992; Glémin 2007). Self-fertilizing populations are thus expected to exhibit low inbreeding depression and high drift load. On the other hand, gBGC, and thus its cost, vanishes as the selfing rate and homozygosity increase (Marais et al. 2004). gBGC could thus challenge classical views on mating systems and it was even speculated that gBGC could affect their evolution (Haudry et al. 2008).Here I present a population genetics model that includes mutation, selection, drift, and gBGC, which extends previous studies (Gutz and Leslie 1976; Lamb and Helmi 1982; Nagylaki 1983a,b; Bengtsson 1990). I specifically examine how gBGC can affect inbreeding depression and the mutation load. I also focus on the effect of mating system, which is especially interesting with regard to the interaction between biased conversion and selection. Finally, I discuss how these results could give insight into how gBGC evolved.

Impacts of gBGC on inbreeding depression:

Inbreeding depression is defined as the reduction in fitness of selfed (and more generally inbred) individuals compared to outcrossed individuals,(15)where and are the mean fitness of outcrosses and selfcrosses, respectively (Charlesworth and Charlesworth 1987; Charlesworth and Willis 2009). The approximation is very good in most conditions, because under weak (s ≪ 1) and strong selection (x ≪ 1) (see Glémin et al. 2003). Similar to the load, considering both sites for which either S or W alleles are deleterious, in proportion q and 1 – q, respectively, we get(16)
gBGC and the genetic basis of inbreeding depression in panmictic populations:
In infinite panmictic populations without gBGC, inbreeding depression depends only on mutation rates and dominance levels. Partially recessive mutations () contribute only to inbreeding depression, and the more recessive they are, the higher the inbreeding depression (Charlesworth and Charlesworth 1987). In finite populations, deterministic results hold for strongly deleterious mutations (s ≫ 1/Ne), which contribute mostly to inbreeding depression. Contrary to the load, weakly deleterious mutations (∼s ≤ 1/Ne) contribute little to inbreeding depression (Figure 4, a and c, and see Bataillon and Kirkpatrick 2000).Open in a separate windowFigure 4.—Inbreeding depression (×106) as a function of s without (a and c) or with (b and d) gBGC (b = 0.0002). (a and b) h = 0.2: thick lines, N = 5000; thin lines, N = 10,000; dashed lines, N = 50,000; dotted lines, N = 100,000. (c and d) N = 10,000: thick lines, h = 0.4; thin lines, h = 0.2; dashed lines, h = 0.1; dotted lines, h = 0.05. u = 10−6, λ = 2.Like the load, gBGC affects both the magnitude and the structure of inbreeding depression. In infinite populations, and more generally for strongly deleterious alleles (Nes ≫ 1), replacing x by xeq given by Equations 4 in Equations 15 and 16 leads to(17a)(17b)(17c)The effect of gBGC on inbreeding depression is not monotonic. Like the load, gBGC increases inbreeding depression if b > hs(1 − 2q/(q + λ − qλ)). However, contrary to the load, a strong gBGC decreases inbreeding depression, which tends to 0 as b increases, while the load tends to qs (Equation 10c). An analysis of Equation 17b shows that mutations that maximize inbreeding depression are those that also maximize the load, i.e., S deleterious mutations with s ≈ 2b.In finite populations, inbreeding depression must be integrated over the Φ distribution, which leads to(18)(see also Glémin et al. 2003). While it is not possible to get an analytical expression of (18), numerical computations (see appendix b) show that S deleterious mutations with s ≈ 2b also maximize inbreeding depression in finite populations (Figure 4). More broadly, inbreeding depression is maximal under the overdominant-like selection regime (gray area in Figure 2). Once again, even low to moderate gBGC markedly affects the genetic structure of inbreeding depression. First, mutations of intermediate effects contribute the most to inbreeding depression, i.e., up to one order of magnitude higher than strongly deleterious mutations (compare Figure 4a with 4b). Second, even nearly additive mutations can have a substantial effect (compare Figure 4c with 4d).Since little is known about the distribution of dominance coefficients, especially the dominance of mildly deleterious mutations (of the order of b), it is difficult to quantitatively predict the full impact of gBGC on inbreeding depression. We can conclude that, on average, gBGC should increase inbreeding depression. However, further insight into mutational parameters is crucial to assess the quantitative impact of gBGC.

Joint effect of gBGC and mating system on the load and inbreeding depression:

Selfing, or more generally inbreeding, slightly reduces the segregating load through the purging of recessive mutations (Ohta and Cockerham 1974), but can substantially increase the fixation load because of the effective population size reduction under inbreeding: (see above and Pollak 1987; Nordborg 1997; Glémin 2007). In numerical examples, I assumed that α decreases with F according to the background selection model (Charlesworth et al. 1993; Nordborg et al. 1996), as in Glémin (2007). With gBGC, selfing thus has two opposite effects on the fixation load. Selfing increases the drift load sensu stricto but decreases the fixation load due to gBGC. A surprising consequence is that the load can be higher in outcrossing than in selfing populations (Figure 5). Quantitatively this is also expected, even with a gBGC hotspot affecting just 3% of the genome (Figure 5 and Open in a separate windowFigure 5.—Effective population size (a and b) and the load (×106) (c–f) as a function of F for different gBGC intensities (thick lines, b = 0; thin lines, b = 0.0001; dashed lines, b = 0.0002; dotted lines, b = 0.0005). The effective population size depends on F under the background selection (BS) model (Charlesworth et al. 1993), using Equations 16 and 17 in Glémin (2007): , where U is the genomic deleterious mutation rate, R is the genomic recombination rate, sd is the mean selection coefficient against strongly deleterious mutations, and hd is their dominance coefficient. N = 10,000, U = 0.2, hd = 0.1, and sd = 0.05. (a, c, and e) R = 5, “weak” BS; (b, d, and f) R = 0.5, “strong” BS. (c and d) Load averaged over half GC and half AT deleterious alleles, with a bias in favor of AT alleles. (e and f) Load averaged over 10% of GC deleterious alleles and 90% of AT deleterious alleles with a bias in favor of AT alleles; see Figure 3. h = 0.5, u = 10−6, and λ = 2.Generally, the effect of selfing is simpler for inbreeding depression. Purging, Ne reduction, and suppression of gBGC contribute to decreasing inbreeding depression in selfing populations (Figure 6a). However, there are special cases in which maximum inbreeding depression is reached for intermediate selfing rates (Figure 6b). In such cases, in outcrossing populations, gBGC is strong enough to sweep polymorphism out and reduce inbreeding depression (b > s, regime 1 in Figure 2). As the selfing rate increases, gBGC declines, and the selection dynamics become overdominant-like (regime 2, Figure 2), thus maximizing inbreeding depression. For high selfing rates, gBGC vanishes (regime 3 in Figure 2) and deleterious alleles are either purged or fixed if there is substantial drift. This is similar to the effect of selfing on inbreeding depression caused by asymmetrical overdominance, where inbreeding depression also peaks for intermediate selfing rates (Ziehe and Roberds 1989; Charlesworth and Charlesworth 1990). In the present case, the range of parameters leading to this peculiar behavior is narrow because the overdominant-like region depends on the selfing rates and can vanish either for low or for high selfing rates (Figure 2).Open in a separate windowFigure 6.—Inbreeding depression (×106) as a function of F for different gBGC intensities (thick lines, b = 0; thin lines, b = 0.0001; dashed lines, b = 0.0002; dotted lines, b = 0.0005). Inbreeding depression is averaged over half GC and half AT deleterious alleles. The effective population size depends on F as in Figure 5 (same parameters). (a) s = 0.002; (b) s = 0.0005; (c) s = 0.0002. h = 0.2, u = 10−6, and λ = 2.

Minimum load and the evolution of gBGC and recombination landscapes:

Although gBGC may have deleterious fitness consequences, it is surprising that it evolved in many taxa (Duret and Galtier 2009a). Birdsell (2002) initially suggested that gBGC may have evolved as a response to mutational bias toward AT (λ > 1, here). Indeed, I show that a minimum load is reached for weak gBGC (b ≈ ln(λ)/4N, Equation 14). This result is very general whatever the distribution of fitness effects of mutations (appendix d). However, the range of optimal gBGC is narrow, and gBGC increases the load as far as b > ln(λ)/2N (appendix c). In humans, using N = 10,000 and λ = 2, gBGC levels that minimize the load are ∼1.17 × 10−5, i.e., one order of magnitude lower than the average bias observed in recombination hotspots (Myers et al. 2005). However, selection on conversion modifiers will not necessarily minimize the load because of gametic disequilibrium generated between modifiers and fitness loci (Bengtsson and Uyenoyama 1990). Selection for limitation of somatic AT-biased mutations could also have selected for GC-biased mismatch repair machinery (Brown and Jiricny 1987). If the bias level that would be selected for somatic reasons is >ln(λ)/2N, a side effect would be the generation of a substantial load at the population level. Finally, it is interesting to note that when synonymous codon positions are under selection for translation accuracy, optimal gBGC levels can be higher than gBGC levels that minimize the protein load, especially when most optimal codons end in G or C ().Conversely, gBGC could also affect the evolution of recombination landscapes, which could evolve to reduce the gBGC load. Surprisingly, for a given recombination/conversion level, the hotspot distribution does not appear to be optimal (Nishant and Rao 2005), one can speculate that the hotspot localization outside genes could be a response to avoid the deleterious effects of gBGC.Up to now, these verbal arguments have not been assessed theoretically (but see Bengtsson and Uyenoyama 1990 for a different kind of conversion bias). Population genetics models are necessary to test these hypotheses concerning the evolution of gBGC and recombination landscapes and to pinpoint the key parameters that might govern their evolution.

gBGC and the evolution of mating systems:

Deleterious mutations also play a crucial role in the evolution of mating systems. They are the main source of inbreeding depression, which balances the automatic advantage of selfing. The drift load is also thought to contribute to the extinction of selfing species. Since they are mainly homozygous, selfing species are mostly free from gBGC and its deleterious impacts. I discuss below how this might affect the evolution of mating systems.
Inbreeding depression and the shift in mating systems:
Inbreeding depression plays a key role in the evolution of mating systems (Charlesworth and Charlesworth 1987; Charlesworth 2006b). Since it balances the automatic advantage of selfing, high inbreeding depression favors outcrossing, while selfing can evolve when it is low. Moreover, selfing helps to purge strongly deleterious mutations, thus decreasing inbreeding depression. This positive feedback reinforces the disruptive selection on the selfing rate and prevents the transition from selfing to outcrossing (Lande and Schemske 1985).Theoretical results suggest that, in most conditions, gBGC would reinforce inbreeding depression in outcrossing populations (Figure 6), which would prevent the evolution of selfing. In reverse, if selfing is initially selected for, recurrent selfing would reduce the load through both purging and avoidance of gBGC. Under this scenario, gBGC would reinforce disruptive selection on mating systems. However, under some conditions (see Figure 6), inbreeding depression peaks at intermediate selfing rates, as observed for asymmetrical overdominance (Ziehe and Roberds 1989; Charlesworth and Charlesworth 1990). In theory, this could prevent the shift toward complete selfing and maintain stable mixed mating systems (Charlesworth and Charlesworth 1990; Uyenoyama and Waller 1991). However, this pattern is observed under restrictive conditions and it is very unlikely on the whole-genome scale. Dominance patterns are crucial for predicting inbreeding depression, especially with gBGC. Contrary to the load, it is thus difficult to evaluate the quantitative impact of gBGC on inbreeding depression. However, increased inbreeding depression in outcrossing species subject to gBGC seems to be the most likely scenario.
gBGC and the long-term evolution of mating systems:
In the long term, the gBGC-induced load also challenges the “dead-end hypothesis,” which posits that, because of the reduction of selection efficacy, self-fertilizing species would accumulate weakly deleterious mutations in the long term, eventually leading to extinction (Takebayashi and Morrell 2001). Because of gBGC, not drift, outcrossing species could also accumulate a load of weakly deleterious mutations (Figure 7), and they could suffer from a higher load than highly self-fertilizing species (Haudry et al. (2008) found that in two outcrossing grass species, but not in two self-fertilizing ones, the dN/dS ratio is significantly higher for genes exhibiting GC enrichment. They speculated that substitutions in these genes might contribute to increasing the load in these two outcrossing grass species. Such results are still very sparse. In plants, evidence of strong gBGC is mainly restricted to grasses (but see Wright et al. 2007). It will be necessary to conduct more in-depth studies to assess the phylogenetic distribution of gBGC in plants and other hermaphrodite organisms and to further test the genomic Achilles'' heel hypothesis in relation to mating systems. While theoretically possible, the quantitative effect of gBGC on the evolution of mating systems remains a new, open, and challenging question.

Conclusion:

I showed that the interaction between gBGC and selection might have surprising qualitative consequences on load and inbreeding depression patterns. Given the few quantitative data available on gBGC levels and selection intensities (mainly in humans), it turns out that even weak genome-wide gBGC can have significant fitness impacts. gBGC should be taken into account not only for sequence analyses (Berglund et al. 2009; Galtier et al. 2009), but also for its potential fitness consequences, for instance concerning genetic diseases. Interferences between gBGC and selection also give rise to new questions on the evolution of mating systems. However, most of the challenging conclusions given here have yet to be quantitatively evaluated. Quantification of gBGC and its interaction with selection in various organisms will be crucial in the future.  相似文献   

7.
Adaptive Divergence in Experimental Populations of Pseudomonas fluorescens. IV. Genetic Constraints Guide Evolutionary Trajectories in a Parallel Adaptive Radiation          下载免费PDF全文
Michael J. McDonald  Stefanie M. Gehrig  Peter L. Meintjes  Xue-Xian Zhang  Paul B. Rainey 《Genetics》2009,183(3):1041-1053
The capacity for phenotypic evolution is dependent upon complex webs of functional interactions that connect genotype and phenotype. Wrinkly spreader (WS) genotypes arise repeatedly during the course of a model Pseudomonas adaptive radiation. Previous work showed that the evolution of WS variation was explained in part by spontaneous mutations in wspF, a component of the Wsp-signaling module, but also drew attention to the existence of unknown mutational causes. Here, we identify two new mutational pathways (Aws and Mws) that allow realization of the WS phenotype: in common with the Wsp module these pathways contain a di-guanylate cyclase-encoding gene subject to negative regulation. Together, mutations in the Wsp, Aws, and Mws regulatory modules account for the spectrum of WS phenotype-generating mutations found among a collection of 26 spontaneously arising WS genotypes obtained from independent adaptive radiations. Despite a large number of potential mutational pathways, the repeated discovery of mutations in a small number of loci (parallel evolution) prompted the construction of an ancestral genotype devoid of known (Wsp, Aws, and Mws) regulatory modules to see whether the types derived from this genotype could converge upon the WS phenotype via a novel route. Such types—with equivalent fitness effects—did emerge, although they took significantly longer to do so. Together our data provide an explanation for why WS evolution follows a limited number of mutational pathways and show how genetic architecture can bias the molecular variation presented to selection.UNDERSTANDING—and importantly, predicting—phenotypic evolution requires knowledge of the factors that affect the translation of mutation into phenotypic variation—the raw material of adaptive evolution. While much is known about mutation rate (e.g., Drake et al. 1998; Hudson et al. 2002), knowledge of the processes affecting the translation of DNA sequence variation into phenotypic variation is minimal.Advances in knowledge on at least two fronts suggest that progress in understanding the rules governing the generation of phenotypic variation is possible (Stern and Orgogozo 2009). The first stems from increased awareness of the genetic architecture underlying specific adaptive phenotypes and recognition of the fact that the capacity for evolutionary change is likely to be constrained by this architecture (Schlichting and Murren 2004; Hansen 2006). The second is the growing number of reports of parallel evolution (e.g., Pigeon et al. 1997; ffrench-Constant et al. 1998; Allender et al. 2003; Colosimo et al. 2004; Zhong et al. 2004; Boughman et al. 2005; Shindo et al. 2005; Kronforst et al. 2006; Woods et al. 2006; Zhang 2006; Bantinaki et al. 2007; McGregor et al. 2007; Ostrowski et al. 2008)—that is, the independent evolution of similar or identical features in two or more lineages—which suggests the possibility that evolution may follow a limited number of pathways (Schluter 1996). Indeed, giving substance to this idea are studies that show that mutations underlying parallel phenotypic evolution are nonrandomly distributed and typically clustered in homologous genes (Stern and Orgogozo 2008).While the nonrandom distribution of mutations during parallel genetic evolution may reflect constraints due to genetic architecture, some have argued that the primary cause is strong selection (e.g., Wichman et al. 1999; Woods et al. 2006). A means of disentangling the roles of population processes (selection) from genetic architecture is necessary for progress (Maynard Smith et al. 1985; Brakefield 2006); also necessary is insight into precisely how genetic architecture might bias the production of mutations presented to selection.Despite their relative simplicity, microbial populations offer opportunities to advance knowledge. The wrinkly spreader (WS) morphotype is one of many different niche specialist genotypes that emerge when experimental populations of Pseudomonas fluorescens are propagated in spatially structured microcosms (Rainey and Travisano 1998). Previous studies defined, via gene inactivation, the essential phenotypic and genetic traits that define a single WS genotype known as LSWS (Spiers et al. 2002, 2003) (Figure 1). LSWS differs from the ancestral SM genotype by a single nonsynonymous nucleotide change in wspF. Functionally (see Figure 2), WspF is a methyl esterase and negative regulator of the WspR di-guanylate cyclase (DGC) (Goymer et al. 2006) that is responsible for the biosynthesis of c-di-GMP (Malone et al. 2007), the allosteric activator of cellulose synthesis enzymes (Ross et al. 1987). The net effect of the wspF mutation is to promote physiological changes that lead to the formation of a microbial mat at the air–liquid interface of static broth microcosms (Rainey and Rainey 2003).Open in a separate windowFigure 1.—Outline of experimental strategy for elucidation of WS-generating mutations and their subsequent identity and distribution among a collection of independently evolved, spontaneously arising WS genotypes. The strategy involves, first, the genetic analysis of a specific WS genotype (e.g., LSWS) to identify the causal mutation, and second, a survey of DNA sequence variation at specific loci known to harbor causal mutations among a collection of spontaneously arising WS genotypes. For example, suppressor analysis of LSWS using a transposon to inactivate genes necessary for expression of the wrinkly morphology delivered a large number of candidate genes (top left) (Spiers et al. 2002). Genetic and functional analysis of these candidate genes (e.g., Goymer et al. 2006) led eventually to the identity of the spontaneous mutation (in wspF) responsible for the evolution of LSWS from the ancestral SM genotype (Bantinaki et al. 2007). Subsequent analysis of the wspF sequence among 26 independent WS genotypes (bottom) showed that 50% harbored spontaneous mutations (of different kinds; see Open in a separate windowFigure 2.—Network diagram of DGC-encoding pathways underpinning the evolution of the WS phenotype and their regulation. Overproduction of c-di-GMP results in overproduction of cellulose and other adhesive factors that determine the WS phenotype. The ancestral SBW25 genome contains 39 putative DGCs, each in principle capable of synthesizing the production of c-di-GMP, and yet WS genotypes arise most commonly as a consequence of mutations in just three DGC-containing pathways: Wsp, Aws, and Mws. In each instance, the causal mutations are most commonly in the negative regulatory component: wspF, awsX, and the phosphodiesterase domain of mwsR (see text).To determine whether spontaneous mutations in wspF are a common cause of the WS phenotype, the nucleotide sequence of this gene was obtained from a collection of 26 spontaneously arising WS genotypes (WSA-Z) taken from 26 independent adaptive radiations, each founded by the same ancestral SM genotype (Figure 1): 13 contained mutations in wspF (Bantinaki et al. 2007). The existence of additional mutational pathways to WS provided the initial motivation for this study.

TABLE 1

Mutational causes of WS
WS genotypeGeneNucleotide changeAmino acid changeSource/reference
LSWSwspFA901CS301RBantinaki et al. (2007)
AWSawsXΔ100-138ΔPDPADLADQRAQAThis study
MWSmwsRG3247AE1083KThis study
WSAwspFT14GI5SBantinaki et al. (2007)
WSBwspFΔ620-674P206Δ (8)aBantinaki et al. (2007)
WSCwspFG823TG275CBantinaki et al. (2007)
WSDwspEA1916GD638GThis study
WSEwspFG658TV220LBantinaki et al. (2007)
WSFwspFC821TT274IBantinaki et al. (2007)
WSGwspFC556TH186YBantinaki et al. (2007)
WSHwspEA2202CK734NThis study
WSIwspEG1915TD638YThis study
WSJwspFΔ865-868R288Δ (3)aBantinaki et al. (2007)
WSKawsOG125TG41VThis study
WSLwspFG482AG161DBantinaki et al. (2007)
WSMawsRC164TS54FThis study
WSNwspFA901CS301RBantinaki et al. (2007)
WSOwspFΔ235-249V79Δ (6)aBantinaki et al. (2007)
WSPawsR222insGCCACCGAA74insATEThis study
WSQmwsR3270insGACGTG1089insDVThis study
WSRmwsRT2183CV272AThis study
WSSawsXC472TQ158STOPThis study
WSTawsXΔ229-261ΔYTDDLIKGTTQThis study
WSUwspFΔ823-824T274Δ (13)aBantinaki et al. (2007)
WSVawsXT74GL24RThis study
WSWwspFΔ149L49Δ (1)aBantinaki et al. (2007)
WSXb???This study
WSYwspFΔ166-180Δ(L51-I55)Bantinaki et al. (2007)
WSZ
mwsR
G3055A
A1018T
This study
Open in a separate windowaP206Δ(8) indicates a frameshift; the number of new residues before a stop codon is reached is in parentheses.bSuppressor analysis implicates the wsp locus (17 transposon insertions were found in this locus). However, repeated sequencing failed to identify a mutation.Here we define and characterize two new mutational routes (Aws and Mws) that together with the Wsp pathway account for the evolution of 26 spontaneously arising WS genotypes. Each pathway offers approximately equal opportunity for WS evolution; nonetheless, additional, less readily realized genetic routes producing WS genotypes with equivalent fitness effects exist. Together our data show that regulatory pathways with specific functionalities and interactions bias the molecular variation presented to selection.  相似文献   

8.
Interaction Between Eye Pigment Genes and Tau-Induced Neurodegeneration in Drosophila melanogaster     
Surendra S. Ambegaokar  George R. Jackson 《Genetics》2010,186(1):435-442
  相似文献   

9.
A Perspective on Micro-Evo-Devo: Progress and Potential     
Maria D. S. Nunes  Saad Arif  Christian Schl?tterer  Alistair P. McGregor 《Genetics》2013,195(3):625-634
  相似文献   

10.
Exciting Prospects for Precise Engineering of Caenorhabditis elegans Genomes with CRISPR/Cas9     
Christian Fr?kj?r-Jensen 《Genetics》2013,195(3):635-642
  相似文献   

11.
Peaks cloaked in the mist: The landscape of mammalian replication origins     
Olivier Hyrien 《The Journal of cell biology》2015,208(2):147-160
  相似文献   

12.
Ion channel regulation by protein S-acylation     
Michael J. Shipston 《The Journal of general physiology》2014,143(6):659-678
Protein S-acylation, the reversible covalent fatty-acid modification of cysteine residues, has emerged as a dynamic posttranslational modification (PTM) that controls the diversity, life cycle, and physiological function of numerous ligand- and voltage-gated ion channels. S-acylation is enzymatically mediated by a diverse family of acyltransferases (zDHHCs) and is reversed by acylthioesterases. However, for most ion channels, the dynamics and subcellular localization at which S-acylation and deacylation cycles occur are not known. S-acylation can control the two fundamental determinants of ion channel function: (1) the number of channels resident in a membrane and (2) the activity of the channel at the membrane. It controls the former by regulating channel trafficking and the latter by controlling channel kinetics and modulation by other PTMs. Ion channel function may be modulated by S-acylation of both pore-forming and regulatory subunits as well as through control of adapter, signaling, and scaffolding proteins in ion channel complexes. Importantly, cross-talk of S-acylation with other PTMs of both cysteine residues by themselves and neighboring sites of phosphorylation is an emerging concept in the control of ion channel physiology. In this review, I discuss the fundamentals of protein S-acylation and the tools available to investigate ion channel S-acylation. The mechanisms and role of S-acylation in controlling diverse stages of the ion channel life cycle and its effect on ion channel function are highlighted. Finally, I discuss future goals and challenges for the field to understand both the mechanistic basis for S-acylation control of ion channels and the functional consequence and implications for understanding the physiological function of ion channel S-acylation in health and disease.Ion channels are modified by the attachment to the channel protein of a wide array of small signaling molecules. These include phosphate groups (phosphorylation), ubiquitin (ubiquitination), small ubiquitin-like modifier (SUMO) proteins (SUMOylation), and various lipids (lipidation). Such PTMs are critical for controlling the physiological function of ion channels through regulation of the number of ion channels resident in the (plasma) membrane; their activity, kinetics, and modulation by other PTMs; or their interaction with other proteins. S-acylation is one of a group of covalent lipid modifications (Resh, 2013). However, unlike N-myristoylation and prenylation (which includes farnesylation and geranylgeranylation), S-acylation is reversible (Fig. 1). Because of the labile thioester bond, S-acylation thus represents a dynamic lipid modification to spatiotemporally control protein function. The most common form of S-acylation, the attachment of the C16 lipid palmitate to proteins (referred to as S-palmitoylation), was first described more than 30 years ago in the transmembrane glycoprotein of the vesicular stomatitis virus and various mammalian membrane proteins (Schmidt and Schlesinger, 1979; Schlesinger et al., 1980). A decade later, S-acylated ion channels—rodent voltage-gated sodium channels (Schmidt and Catterall, 1987) and the M2 ion channel from the influenza virus (Sugrue et al., 1990)—were first characterized. Since then, more than 50 distinct ion channel subunits have been experimentally demonstrated to be S-acylated (El-Husseini and Bredt, 2002; Linder and Deschenes, 2007; Fukata and Fukata, 2010; Greaves and Chamberlain, 2011; Resh, 2012). In the last few years, with the cloning of enzymes controlling S-acylation and development of various proteomic tools, we have begun to gain substantial mechanistic and physiological insight into how S-acylation may control multiple facets of the life cycle of ion channels: from their assembly, through their trafficking and regulation at the plasma membrane, to their final degradation (Fig. 2).Open in a separate windowFigure 1.Protein S-acylation: a reversible lipid posttranslational modification of proteins. (A) Major lipid modifications of proteins. S-acylation is reversible due to the labile thioester bond between the lipid (typically, but not exclusively, palmitate) and the cysteine amino acid of is target protein. Other lipid modifications result from stable bond formation between either the N-terminal amino acid (amide) or the amino acid side chain in the protein (thioether and oxyester). The zDHHC family of palmitoyl acyltransferases mediates S-acylation with other enzyme families controlling other lipid modifications: N-methyltransferase (NMT) controls myristoylation of many proteins such as the src family kinase, Fyn kinase; and amide-linked palmitoylation of the secreted sonic hedgehog protein is mediated by Hedgehog acyltransferase (Hhat), a membrane-bound O-acyl transferase (MBOAT) family. Prenyl transferases catalyze farnesyl (farnesyltransferase, FTase) or geranylgeranyl (geranylgeranyl transferase I [GGTase I] and geranylgeranyl transferase II [GGTase II]) in small GTPase proteins such as RAS and the Rab proteins, respectively. Porcupine (Porcn) is a member of the MBOAT family acylates secreted proteins such as Wnt. (B) zDHHC enzymes typically use coenzyme A (CoA)-palmitate; however, other long chain fatty acids (either saturated or desaturated) can also be used. Deacylation is mediated by several acylthioesterases of the serine hydrolase family. (C) zDHHC acyltransferases (23 in humans) are predicted transmembrane proteins (typically with 4 or 6 transmembrane domains) with the catalytic DHHC domain located in a cytosolic loop.

Table 1.

Pore-forming subunits of ion channels experimentally determined to be S-acylated
ChannelSubunitGeneCandidate S-acylation sitesUniProt IDReferences
Ligand-gated
AMPAGluA1Gria1593FSLGAFMQQGCDISPRSLSGRIP23818Hayashi et al., 2005
819LAMLVALIEFCYKSRSESKRMKP23818Hayashi et al., 2005
GluA2Gria2600FSLGAFMRQGCDISPRSLSGRIP23819Hayashi et al., 2005
826LAMLVALIEFCYKSRAEAKRMKP23819Hayashi et al., 2005
GluA3Gria3605FSLGAFMQQGCDISPRSLSGRIQ9Z2W9Hayashi et al., 2005
831LAMMVALIEFCYKSRAESKRMKQ9Z2W9Hayashi et al., 2005
GluA4Gria4601FSLGAFMQQGCDISPRSLSGRIQ9Z2W8Hayashi et al., 2005
827LAMLVALIEFCYKSRAEAKRMKQ9Z2W8Hayashi et al., 2005
GABAAγ2Gabrg2405QERDEEYGYECLDGKDCASFFCCFEDCRTGAWRHGRIP22723Rathenberg et al., 2004; Fang et al., 2006
KainateGluK2Grik2848KNAQLEKRSFCSAMVEELRMSLKCQRRLKHKPQAPVP39087Pickering et al., 1995
nAChRα4Chrna4263TVLVFYLPSECGEKVTLCISVO70174Alexander et al., 2010; Amici et al., 2012
α7Chrna7NDAlexander et al., 2010; Drisdel et al., 2004
β2Chrnb2NDAlexander et al., 2010
NMDAGluN2AGrin2a838EHLFYWKLRFCFTGVCSDRPGLLFSISRGIYSCIHGVHIEEKKP35436Hayashi et al., 2009
1204SDRYRQNSTHCRSCLSNLPTYSGHFTMRSPFKCDACLRMGNLYDIDP35436Hayashi et al., 2009
GluN2BGrin2b839EHLFYWQFRHCFMGVCSGKPGMVFSISRGIYSCIHGVAIEERQQ01097Hayashi et al., 2009
1205DWEDRSGGNFCRSCPSKLHNYSSTVAGQNSGRQACIRCEACKKAGNLYDISQ01097Hayashi et al., 2009
P2X7P2X7P2rx7361AFCRSGVYPYCKCCEPCTVNEYYYRKKQ9Z1M0Gonnord et al., 2009
469APKSGDSPSWCQCGNCLPSRLPEQRRQ9Z1M0Gonnord et al., 2009
488PEQRRALEELCCRRKPGRCITTQ9Z1M0Gonnord et al., 2009
562DMADFAILPSCCRWRIRKEFPKQ9Z1M0Gonnord et al., 2009
Voltage gated
Potassium
BK, maxiKKCa1.1Kcnma143WRTLKYLWTVCCHCGGKTKEAQKIQ08460Jeffries et al., 2010
635MSIYKRMRRACCFDCGRSERDCSCMQ08460Tian et al., 2008; 2010
Kv1.1Kcna1233SFELVVRFFACPSKTDFFKNIP16388Gubitosi-Klug et al., 2005
Kv1.5Kcna516LRGGGEAGASCVQSPRGECGCQ61762Jindal et al., 2008
583VDLRRSLYALCLDTSRETDL-stopQ61762Zhang et al., 2007; Jindal et al., 2008
SodiumNaV1.2Scn2a1NDSchmidt and Catterall, 1987
640MNGKMHSAVDCNGVVSLVGGPP04775Bosmans et al., 2011
1042LEDLNNKKDSCISNHTTIEIGP04775Bosmans et al., 2011
1172TEDCVRKFKCCQISIEEGKGKP04775Bosmans et al., 2011
Other channels
AquaporinAQP4Aqp43DRAAARRWGKCGHSCSRESIMVAFKP55088Crane and Verkman, 2009; Suzuki et al., 2008
CFTRCFTRCFTR514EYRYRSVIKACQLEEDISKFAEKDP13569McClure et al., 2012
1385RRTLKQAFADCTVILCEHRIEAP13569McClure et al., 2012
ConnexinCx32Gjb1270GAGLAEKSDRCSAC-stopP28230Locke et al., 2006
ENaCENaC βScnn1b33TNTHGPKRIICEGPKKKAMWFLQ9WU38Mueller et al., 2010
547WITIIKLVASCKGLRRRRPQAPYQ9WU38Mueller et al., 2010
ENaC γScnn1g23PTIKDLMHWYCLNTNTHGCRRIVVSRGRLQ9WU39Mukherjee et al., 2014
Influenza M2M240LWILDRLFFKCIYRFFEHGLKQ20MD5Sugrue et al., 1990; Holsinger et al., 1995; Veit et al., 1991
RyR1RYR1Ryr114LRTDDEVVLQCSATVLKEQLKLCLAAEGFGNRLP11716Chaube et al., 2014
110RHAHSRMYLSCLTTSRSMTDKP11716Chaube et al., 2014
243RLVYYEGGAVCTHARSLWRLEP11716Chaube et al., 2014
295EDQGLVVVDACKAHTKATSFCP11716Chaube et al., 2014
527ASLIRGNRANCALFSTNLDWVP11716Chaube et al., 2014
1030ATKRSNRDSLCQAVRTLLGYGP11716Chaube et al., 2014
1664SHTLRLYRAVCALGNNRVAHAP11716Chaube et al., 2014
2011HFKDEADEEDCPLPEDIRQDLP11716Chaube et al., 2014
2227KMVTSCCRFLCYFCRISRQNQP11716Chaube et al., 2014
2316KGYPDIGWNPCGGERYLDFLRP11716Chaube et al., 2014
2353VVRLLIRKPECFGPALRGEGGP11716Chaube et al., 2014
2545EMALALNRYLCLAVLPLITKCAPLFAGTEHRP11716Chaube et al., 2014
3160DVQVSCYRTLCSIYSLGTTKNTYVEKLRPALGECLARLAAAMPVP11716Chaube et al., 2014
3392LLVRDEFSVLCRDLYALYPLLP11716Chaube et al., 2014
3625SKQRRRAVVACFRMTPLYNLPP11716Chaube et al., 2014
Open in a separate windowCommon channel abbreviation and subunit as well as gene names are given. Candidate S-acylation sites: experimentally determined cysteine residues (bold) with flanking 10 amino acids. Underlines indicate predicted transmembrane domains. Amino acid numbering corresponds to the UniProt ID. References: selected original supporting citations.Open in a separate windowFigure 2.Protein S-acylation and regulation of the ion channel lifecycle zDHHCs are found in multiple membrane compartments and regulate multiple steps in the ion channel lifecycle including: (1) assembly and (2) ER exit; (3) maturation and Golgi exit; (4) sorting and trafficking; (5) trafficking and insertion into target membrane; (6) clustering and localization in membrane microdomains; control of properties, activity (7), and regulation by other signaling pathways; and (8) internalization, recycling, and final degradation.

Table 3.

Other channels identified in mammalian palmitoylome screens
ChannelGene
Anion
Chloride channel 6Clcn6
Chloride intracellular channel 1Clic1
Chloride intracellular channel 4Clic4
Tweety homologue 1Ttyh1
Tweety homologue 3Ttyh3
Voltage-dependent anion channel 1Vdac1
Voltage-dependent anion channel 2Vdac2
Voltage-dependent anion channel 3Vdac3
Calcium
Voltage-dependent, L-type subunit α 1SCacna1s
Voltage-dependent, gamma subunit 8Cacng8
Cation
Amiloride-sensitive cation channel 2Accn2
Glutamate
Ionotropic, Δ1Grid1
Perforin
Perforin 1Prf1
Potassium
Voltage-gated channel, subfamily Q, member 2Kcnq2
Sodium
Voltage-gated, type I, αScn1a
Voltage-gated, type III, αScn3a
Voltage-gated, type IX, αScn9a
Transient receptor potential
Cation channel, subfamily V, member 2Trpv2
Cation channel, subfamily M, member 7Trpm7
Open in a separate windowChannels identified in global S-acylation screens (Wan et al., 2007, 2013; Kang et al., 2008; Martin and Cravatt, 2009; Yang et al., 2010; Yount et al., 2010; Merrick et al., 2011; Wilson et al., 2011; Jones et al., 2012; Ren et al., 2013; Chaube et al., 2014) and not independently characterized as in and2.2. Common channel abbreviation and gene names are given.Here, I provide a primer on the fundamentals of S-acylation, in the context of ion channel regulation, along with a brief overview of tools available to interrogate ion channel S-acylation. I will discuss key examples of how S-acylation controls distinct stages of the ion channel life cycle before highlighting some of the key challenges for the field in the future.

Fundamentals of S-acylation: The what, when, where, and how

S-acylation: A fatty modification that controls multiple aspects of protein function.

Protein S-acylation results from the attachment of a fatty acid to intracellular cysteine residues of proteins via a labile, thioester linkage (Fig. 1, A and B). Because the thioester bond is subject to nucleophilic attack, S-acylation, unlike other lipid modifications such as N-myristoylation and prenylation, is reversible. However, for most ion channels, as for other S-acylated proteins, the dynamics of S-acylation are poorly understood. Distinct classes of proteins can undergo cycles of acylation and deacylation that are very rapid (e.g., on the timescale of seconds, as exemplified by rat sarcoma [RAS] proteins), much longer (hours), or essentially irreversible during the lifespan of the protein (El-Husseini and Bredt, 2002; Linder and Deschenes, 2007; Zeidman et al., 2009; Fukata and Fukata, 2010; Greaves and Chamberlain, 2011; Resh, 2012). For most ion channels, in fact most S-acylated proteins, the identity of the native lipid species attached to specific cysteine residues is also largely unknown. However, the saturated C16:0 lipid palmitate is commonly thought to be the major lipid species in many S-acylated proteins (Fig. 1). Indeed, much of the earliest work on S-acylation involved the metabolic labeling of proteins in cells with tritiated [3H]palmitate, an approach that still remains useful and important. However, lipids with different chain lengths and degrees of unsaturation (such as oleic and stearic acids) can also be added to cysteines via a thioester linkage, potentially allowing differential control of protein properties through the attachment of distinct fatty acids (El-Husseini and Bredt, 2002; Linder and Deschenes, 2007; Zeidman et al., 2009; Fukata and Fukata, 2010; Greaves and Chamberlain, 2011; Resh, 2012).S-acylation increases protein hydrophobicity and has thus been implicated in controlling protein function in many different ways. Most commonly, as with membrane-associated proteins like RAS and postsynaptic density protein 95 (PSD-95), S-acylation controls membrane attachment and intracellular trafficking. However, S-acylation can also control protein–protein interactions, protein targeting to membrane subdomains, protein stability, and regulation by other PTMs such as phosphorylation (El-Husseini and Bredt, 2002; Fukata and Fukata, 2010; Linder and Deschenes, 2007; Greaves and Chamberlain, 2011; Shipston, 2011; Resh, 2012). Evidence for all these mechanisms in controlling ion channel function is beginning to emerge.

Enzymatic control of S-acylation by zinc finger–containing acyltransferase (zDHHC) transmembrane acyltransferases.

Although autoacylation of some proteins has been reported in the presence of acyl coenzyme A (acyl-CoA; Linder and Deschenes, 2007), most cellular S-acylation, in organisms from yeast to humans, is thought to be enzymatically driven by a family of protein acyltransferases (gene family: zDHHC, with ∼23 members in mammals). These acyltransferases are predicted to be transmembrane zinc finger containing proteins (Fig. 1 C) that include a conserved Asp-His-His-Cys (DHHC) signature sequence within a cysteine-rich stretch of ∼50 amino acids critical for catalytic activity (Fukata et al., 2004). Although the enzymatic activity and lipid specificity of all of the zDHHC family proteins has not been elucidated, S-acylation is thought to proceed through a common, two step “ping pong” process (Mitchell et al., 2010; Jennings and Linder, 2012). However, different zDHHC enzymes may show different acyl-CoA substrate specificities. For example, zDHHC3 activity is reduced by acyl chains of >16 carbons (e.g., stearoyl CoA), whereas zDHHC2 efficiently transfers acyl chains of 14 carbons or longer (Jennings and Linder, 2012). The local availability of different acyl-CoA species may thus play an important role in differentially controlling protein S-acylation.We know very little about how zDHHC activity and function are regulated. Dimerization of zDHHCs 2 and 3 reduces their zDHHC activity compared with the monomeric form (Lai and Linder, 2013). Moreover, zDHHCs undergo autoacylation and contain predicted sites for other posttranslational modifications. Almost half of all mammalian zDHHCs contain a C-terminal PSD-95, Discs large, and ZO-1 (PDZ) domain binding motif, allowing them to assemble with various PDZ domain proteins that regulate ion channels (such as GRIP1b and PSD-95; Thomas and Hayashi, 2013). Other protein interaction domains are also observed in zDHHCs, such as ankyrin repeats in zDHHC17 and zDHHC13 (Greaves and Chamberlain, 2011). Indeed, increasing evidence suggests that various ion channels—including the ligand-gated γ-aminobutyric (GABAA), α-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate (AMPA), and NMDA receptors and the large conductance calcium- and voltage-activated (BK) potassium channels—can assemble in complexes with their cognate zDHHCs.The expansion of the number of zDHHCs in mammals (23 vs. 7 in yeast), together with increased prevalence of PDZ interaction motifs, likely represents evolutionary gain-of-function mechanisms to diversify zDHHC function (Thomas and Hayashi, 2013). Evolutionary gain of function is also seen in ion channel subunit orthologues through acquisition of S-acylated cysteine residues absent in orthologues lower in the phylogenetic tree (such as the transmembrane domain 4 [TM4] sites in GluA1–4 subunits of AMPA receptors [Thomas and Hayashi, 2013] and the sites in the alternatively spliced stress-regulated exon [STREX] insert in the C terminus of the BK channel [Tian et al., 2008]). Importantly, some zDHHCs may have additional roles beyond their acyltransferase function. For example, the Drosophila melanogaster zDHHC23 orthologue lacks the catalytic DHHC sequence, and thus protein acyltransferase activity, and is a chaperone involved in protein trafficking (Johswich et al., 2009), whereas mammalian zDHHC 23 has a functional zDHHC motif and, in addition to S-acylating BK channels (Tian et al., 2012), can bind and regulate, but does not S-acylate, neuronal nitric oxide synthase (nNOS; Saitoh et al., 2004).However, as with most S-acylated proteins, the identity of the zDHHCs that modify specific cysteine residues on individual ion channels is not known. Indeed, relatively few studies have tried to systematically identify the zDHHCs controlling ion channel function (Tian et al., 2010, 2012). Thus we are largely ignorant of the extent to which different zDHHCs may have specific ion channel targets or may display specificity. Some details are beginning to emerge: for example, zDHHC3 appears to be a rather promiscuous acyltransferase reported to S-acylate several ion channels (Keller et al., 2004; Hayashi et al., 2005, 2009; Tian et al., 2010), whereas distinct sites on the same ion channel subunit can be modified by distinct subsets of zDHHCs (Tian et al., 2010, 2012). Although we are still in the foothills of understanding the substrates and physiological roles of different zDHHCs, mutation or loss of function in zDHHCs is associated with an increasing number of human disorders, including cancers, various neurological disorders (such as Huntington’s disease and X-linked mental retardations), and disruption of endocrine function in diabetes (Linder and Deschenes, 2007; Fukata and Fukata, 2010; Greaves and Chamberlain, 2011; Resh, 2012).

Deacylation is controlled by acylthioesterases.

Protein deacylation is enzymatically driven by a family of acylthioesterases that belong to the serine hydrolase superfamily (Zeidman et al., 2009; Bachovchin et al., 2010). Indeed, using a broad spectrum serine lipase inhibitor, global proteomic S-acylation profiling identified a subset of serine hydrolases responsible for depalmitoylation (Martin et al., 2012). This study identified both the previously known acylthioesterases as well as potential novel candidate acylthioesterases. The acylthioesterases responsible for deacylating ion channels, as for most other acylated membrane proteins, have not been clearly defined. Furthermore, the extent to which different members of the serine hydrolase superfamily display acylthioesterase activity toward ion channels is not known. Moreover, whether additional mechanisms of nucleophilic attack of the labile thioester bond may also mediate deacylation is not known.Homeostatic control of deacylation of many signaling proteins is likely affected by a family of cytosolic acyl protein thioesterases including lysophospholipase 1 (LYPLA1; Yeh et al., 1999; Devedjiev et al., 2000) and lysophospholipase 2 (LYPLA2; Tomatis et al., 2010). These enzymes show some selectivity for different S-acylated peptides (Tomatis et al., 2010). Indeed, LYPLA1, but not LYPLA2, deacylates the S0-S1 loop of BK channels, leading to Golgi retention of the channel (Tian et al., 2012). A splice variant of the related LYPLAL1 acylthioesterases can also deacylate the BK channel S0-S1 loop, although the crystal structure of LYPLAL1 suggests it is likely to have a preference for lipids with shorter chains than palmitate (Bürger et al., 2012). Thus, whether lipid preference depends on protein interactions or if BK channels have multiple lipid species at the multicysteine S0-S1 site remain unknown. Relatively little is known about the regulation of these acylthioesterases; however, both LYPLA1 and LYPLA2 are themselves S-acylated. This controls their trafficking and association with membranes (Kong et al., 2013; Vartak et al., 2014) and may be important for accessing the thioesterase bond at the membrane interface. Additional mechanisms may promote accessibility of thioesterases to target cysteines. For example, the prolyl isomerase protein FKBP12 binds to palmitoylated RAS, and promotes RAS deacylation via a proline residue near the S-acylated cysteine (Ahearn et al., 2011).Upon lysosomal degradation, many proteins are deacylated by the lysosomal palmitoyl protein thioesterase (PPT1; Verkruyse and Hofmann, 1996), and mutations in PPT1 lead to the devastating condition of infantile neuronal ceroid lipofuscinosis (Vesa et al., 1995; Sarkar et al., 2013). However, PPT1 can also be found in synaptic and other transport vesicles, and genetic deletion of PPT1 in mice may have different effects on similar proteins, which suggests roles beyond just lysosomal mediated degradation. For example, in PPT1 knockout mice the total expression and surface membrane abundance of the GluA4 AMPA receptor subunit was decreased, whereas PPT1 knockout had no effect on GluA1 or GluA2 AMPA subunits nor on NMDA receptor subunit expression or surface abundance (Finn et al., 2012).However, for most ion channels, the questions of which enzymes control deacylation, where this occurs in cells, and how the time course of acylation–deacylation cycles are regulated are largely unknown. Thus, whether deacylation plays an active role in channel regulation remains poorly understood.

S-acylation occurs at membrane interfaces.

Because the zDHHCs are transmembrane proteins and the catalytic DHHC domain is located at the cytosolic interface with membranes (Fig. 1 C), S-acylation of ion channels occurs at membrane interfaces. Although overexpression studies of recombinant mammalian zDHHCs in heterologous expression systems have indicated that most zDHHCs are localized to either the endoplasmic reticular or Golgi apparatus membranes (or both; Ohno et al., 2006), some zDHHCs are also found in other compartments, including the plasma membrane and trafficking endosomes (Thomas et al., 2012; Fukata et al., 2013). We know very little about the regulation and subcellular localization of most native zDHHC enzymes in different cell types, in large part because of the lack of high-quality antibodies that recognize native zDHHCs. However, some enzymes, including zDHHC2, can dynamically shuttle between different membrane compartments. Activity-dependent redistribution of zDHHC2 in neurons (Noritake et al., 2009) controls S-acylation of the postsynaptic scaffolding protein PSD-95, thereby regulating NMDA receptor function. Intriguingly, as ion channels themselves determine cellular excitability, this may provide a local feedback mechanism to regulate S-acylation status. Thus, although different zDHHCs may reside in multiple membrane compartments through which ion channels traffic, the subcellular location at which most ion channels are S-acylated, as well as the temporal dynamics, is largely unknown. As discussed below (see the “Tools to analyze ion channel S-acylation” section), we are starting to unravel some of the details, with ER exit, Golgi retention, recycling endosomes, and local plasma membrane compartments being key sites in the control of ion channel S-acylation (Fig. 2).

Local membrane and protein environment determines cysteine S-acylation.

The efficiency of S-acylation of cysteine residues is likely enhanced by its localization at membranes because the local concentration of fatty acyl CoA is increased near hydrophobic environments (Bélanger et al., 2001). Furthermore, S-acylation of polytopic transmembrane proteins such as ion channels would be facilitated when S-acylated cysteines are bought into close proximity of membranes by membrane targeting mechanisms such as transmembrane helices (Figs. 3 and and4).4). However, the S-acylated cysteine is located within 10 amino acids of a transmembrane domain in only ∼20% of identified S-acylated ion channel subunits, such as the TM4 site of GluA1–4 (and2).2). Most S-acylated cysteines are located either within intracellular loops (∼40%: Fig. 3, A and B) or the N- or C-terminal cytosolic domains (∼5% and 35%, respectively; Fig. 3, A and B). Furthermore, the majority of S-acylated cysteines located in intracellular loops or intracellular N- or C-terminal domains of ion channel subunits are within predicted regions of protein disorder (Fig. 3 B). This suggests that S-acylation may provide a signal to promote conformational restraints on such domains, in particular by providing a membrane anchor. For these sites, additional initiating membrane association signals are likely required adjacent to the site of S-acylation. Likely candidates include other hydrophobic domains (as for the TM2 site in GluA1–4 subunits; Fig. 4 A) and other lipid anchors (e.g., myristoylation in src family kinases, such as Fyn kinase). However, in >30% of S-acylated ion channels, the S-acylated cysteine is juxtaposed to a (poly) basic region of amino acids that likely allows electrostatic interaction with negative membrane phospholipids. The BK channel pore-forming α subunit, encoded by the KCNMA1 gene, provides a clear example of this latter mechanism. This channel is S-acylated within an alternatively spliced domain (STREX) in its large intracellular C terminus (Fig. 4 C). Immediately upstream of the S-acylated dicysteine motif is a polybasic region enriched with arginine and lysine. Site-directed mutation of these basic amino acids disrupts S-acylation of the downstream cysteine residues (Jeffries et al., 2012). Furthermore, phosphorylation of a consensus PKA site (i.e., introduction of negatively charged phosphate) into the polybasic domain prevents STREX S-acylation. Thus, at the STREX domain, an electrostatic switch, controlled by phosphorylation, is an important determinant of BK channel S-acylation. In other proteins, cysteine reactivity is also enhanced by proximity to basic (or hydrophobic) residues (Bélanger et al., 2001; Britto et al., 2002; Kümmel et al., 2010). Furthermore, cysteine residues are subject to a range of modifications including nitrosylation, sulphydration, reduction-oxidation (REDOX) modification, and formation of disulphide bonds (Sen and Snyder, 2010). Evidence is beginning to emerge that these reversible modifications are mutually competitive for S-acylation of target cysteines (see the “S-acylation and posttranslational cross-talk controls channel trafficking and activity” section; Ho et al., 2011; Burgoyne et al., 2012).Open in a separate windowFigure 3.S-acylation sites in ion channel pore-forming subunits. (A) Schematic illustrating different locations of cysteine S-acylation in transmembrane ion channels subunits. (B) Relative proportion of identified S-acylated cysteine residues: in each location indicated in A (top); in -C-, -CC-, or -Cx(2–3)C- motifs (middle); or in cytosolic regions of predicted protein disorder (bottom; determined using multiple algorithms on the DisProt server, http://www.disprot.org/metapredictor.php; Sickmeier et al., 2007) for transmembrane ion channel pore-forming subunits.Open in a separate windowFigure 4.Multisite S-acylation in ion channels controls distinct functions. (A–C) Schematic illustrating location of multiple S-acylated domains in AMPA receptor GluA1–4 subunits (A), NMDA receptor GluN2A subunits (B), and BK channel pore-forming α subunits (C), encoded by the Kcnma1 gene. Each domain confers distinct functions/properties on the respective ion channel and is regulated by distinct zDHHCs (see the “Control of ion channel cell surface expression and spatial organization in membranes” section for further details).

Table 2.

Accessory subunits and selected ion channel adapter proteins
ChannelSubunitGeneCandidate S-acylation sitesUniProt IDReferences
Voltage gated
CalciumCaVβ2aCacnb21MQCCGLVHRRRVRVQ8CC27Chien et al., 1996; Stephens et al., 2000; Heneghan et al., 2009; Mitra-Ganguli et al., 2009
PotassiumKChip2Kcnip234LKQRFLKLLPCCGPQALPSVSEQ9JJ69Takimoto et al., 2002
KChip3Kcnip335PRFTRQALMRCCLIKWILSSAAQ9QXT8Takimoto et al., 2002
BK β4Kcnmb4193VGVLIVVLTICAKSLAVKAEAQ9JIN6Chen et al., 2013
Adapter proteins that interact with ion channelsPICK1Pick1404TGPTDKGGSWCDS-stopQ62083Thomas et al., 2013
Grip1bGrip11MPGWKKNIPICLQAEEQEREQ925T6-2Thomas et al., 2012; Yamazaki et al., 2001
psd-95Dlg41MDCLCIVTTKKYRQ62108Topinka and Bredt, 1998
S-delphilinGrid2ip1MSCLGIFIPKKHQ0QWG9-2Matsuda et al., 2006
Ankyrin-GAnk360YIKNGVDVNICNQNGLNALHLF1LNM3He et al., 2012
Open in a separate windowCommon channel abbreviation and subunit as well as gene names are given. Candidate S-acylation sites: experimentally determined cysteine residues (bold) with flanking 10 amino acids. Underlines indicate predicted transmembrane domains. Amino acid numbering corresponds to the UniProt ID. References: selected original supporting citations.Although these linear amino acid sequence features are likely to be important for efficient S-acylation, there is no canonical “consensus” S-acylation motif analogous to the linear amino acid sequences that predict sites of phosphorylation. Of the experimentally validated ion channel subunits shown to be S-acylated, ∼70% of candidate S-acylated cysteines are predominantly characterized as single cysteine (-C-) motifs, whereas dicysteine motifs (-CC-) and (CX(1–3)C-) motifs comprise ∼10% and 20% of all sites, respectively (Fig. 3 B). However, several freely available online predictive tools have proved successful in characterizing potential new palmitoylation targets. In particular, the latest iteration of the multiplatform CSS-palm 4.0 tool (Ren et al., 2008) exploits a Group-based prediction algorithm by comparing the surrounding amino acid sequence similarity to that of a set of 583 experimentally determined S-acylation sites from 277 distinct proteins. CSS-palm 4.0 predicts >80% of the experimentally identified ion channel S-acylation sites (Location of S-acylated cysteine is important for differential control of channel function.Many proteins are S-acylated at multiple sites. A remarkable example of this, in the ion channel field, is the recent identification of 18 S-acylated cysteine residues in the skeletal muscle ryanodine receptor/Ca2+-release channel (RyR1). The S-acylated cysteine residues are distributed throughout the cytosolic N terminus, including domains important for protein–protein interactions (Chaube et al., 2014). Although deacylation of skeletal muscle RyR1 reduces RyR1 activity, the question of which of these cysteine residues in RyR1 are important for this effect and whether distinct S-acylated cysteines in RyR1 control different functions and/or properties remains to be determined.However, both ligand-gated (NMDA and AMPA) and voltage-gated (BK) channels provide remarkable insights into how S-acylation of different domains within the same polytopic protein can exert fundamentally distinct effects (Fig. 4). For example, S-acylation of the hydrophobic cytosolic TM2 domain located at the membrane interface of the AMPA GluA1 subunit (Fig. 4 A) decreases AMPA receptor surface expression by retaining the subunit at the Golgi apparatus (Hayashi et al., 2005). In contrast, depalmitoylation of the C-terminal cysteine in GluA1 results in enhanced PKC-dependent phosphorylation of neighboring serine residues, which results in increased interaction with the actin-binding protein 4.1N in neurons, leading to enhanced AMPA plasma membrane insertion (Lin et al., 2009). S-acylation of the C-terminal cluster of cysteine residues (Fig. 4 B, Cys II site) in GluN2A and GluN2B controls Golgi retention, whereas palmitoylation of the cysteine cluster (Cys I site) proximal to the M4 transmembrane domain controls channel internalization (Hayashi et al., 2009). Distinct roles of S-acylation on channel trafficking and regulation are also observed in BK channels (Figs. 4 C and and5).5). S-acylation of the N-terminal intracellular S0-S1 linker controls surface expression, in part by controlling ER and Golgi exit of the channel (Jeffries et al., 2010; Tian et al., 2012), whereas S-acylation of the large intracellular C terminus, within the alternatively spliced STREX domain, controls BK channel regulation by AGC family protein kinases (Tian et al., 2008; Zhou et al., 2012).Open in a separate windowFigure 5.S-acylation controls BK channel trafficking and regulation by AGC family protein kinases via distinct sites. The BK channel STREX splice variant pore-forming α subunit is S-acylated at two sites: the S0-S1 loop and the STREX domain in the large intracellular C terminus. S-acylation of the S0-S1 loop promotes high surface membrane expression of the channel; thus, deacylation of this site decreases the number of channels at the cell surface (see the “Control of ion channel cell surface expression and spatial organization in membranes” section for further details). In contrast, S-acylation of the STREX domain allows inhibition of channel activity by PKA-mediated phosphorylation of a PKA serine motif (closed hexagon) immediately upstream of the palmitoylated cysteine residues in STREX. In the S-acylated state, PKC has no effect on channel activity even though a PKC phosphorylation site serine motif is located immediately downstream of the STREX domain (open triangle). Deacylation of STREX dissociates the STREX domain from the plasma membrane, and exposes the PKC serine motif so that it can now be phosphorylated by PKC (closed triangle), resulting in channel inhibition. In the deacylated state, PKA has no effect on channel activity (open hexagon). Thus, deacylation of the STREX domain switches channel regulation from a PKA-inhibited to a PKC-inhibited phenotype (see the “S-acylation and posttranslational cross-talk controls channel trafficking and activity” section for further details).How does S-acylation of distinct domains control such behavior, and are distinct sites on the same protein acylated by distinct zDHHCs? A systematic small interfering RNA (siRNA) screen of zDHHC enzymes mediating BK channel S-acylation indicated that distinct subsets of zDHHCs modify discrete sites. The S0-S1 loop is S-acylated by zDHHCs 22 and 23, whereas the STREX domain is S-acylated by several zDHHCs including 3, 9, and 17 (Tian et al., 2008, 2012). In both cases, each domain has two distinct S-acylated cysteines; however, whether these cysteines are differentially S-acylated by specific zDHHCs is unknown, Furthermore, whether multiple zDHHCs are required because the domains undergo repeated cycles of S-acylation and deacylation, and thus different zDHHCs function at different stages of the protein lifecycle, remains to be determined. Although systematic siRNA screens have, to date, not been performed on other ion channels, data from other multiply S-acylated channels, such as NMDA, AMPA, and BK channel subunits, supports the hypothesis that zDHHCs can show substrate specificity (Hayashi et al., 2005, 2009; Tian et al., 2010).It is generally assumed that S-acylation facilitates the membrane association of protein domains. This is clearly the case for peripheral membrane proteins, such as RAS or PSD-95, but direct experimental evidence for S-acylation controlling membrane association of the cytosolic domains of transmembrane proteins is largely elusive. One of the best examples involves the large C-terminal domain of the BK channel, which comprises more than two-thirds of the pore-forming subunit (Fig. 5). In the absence of S-acylation of the STREX domain, or exclusion of the 59–amino acid STREX insert, the BK channel C terminus is cytosolic (Tian et al., 2008). However, if the STREX domain is S-acylated, the entire C terminus associates with the plasma membrane, a process that can be dynamically regulated by phosphorylation of a serine immediately upstream of the S-acylated cysteines in the STREX domain (Tian et al., 2008). This S-acylation–dependent membrane association markedly affects the properties and regulation of the channel (Jeffries et al., 2012) and has been proposed to confer significant structural rearrangements. In support of such structural rearrangement, S-acylated STREX channels are not inhibited by PKC-dependent phosphorylation even though a PKC phosphorylation site serine motif, conserved in other BK channel variants, is present downstream of the STREX domain. In other BK channel variants lacking the STREX insert, this PKC site is required for channel inhibition by PKC-dependent phosphorylation. However, after deacylation of the STREX domain, PKC can now phosphorylate this PKC phosphorylation serine motif, which suggests that the site has become accessible, consequently resulting in channel inhibition (Fig. 5; Zhou et al., 2012).How might S-acylation of a cysteine residue juxtaposed to another membrane anchoring domain control protein function? The simplest mechanism would involve acting as an additional anchor (Fig. 3 A). In some systems, juxta-transmembrane palmitoylation allows tilting of transmembrane domains, effectively shortening the transmembrane domain to reduce hydrophobic mismatch (Nyholm et al., 2007), particularly at the thinner ER membrane (Abrami et al., 2008; Charollais and Van Der Goot, 2009; Baekkeskov and Kanaani, 2009), and confer conformational restraints on the peptide (Fig. 3 A). Such a mechanism has been proposed to control ER exit of the regulatory β4 subunits of BK channels. In this case, depalmitoylation of a cysteine residue juxtaposed to the second transmembrane domain of the β4 subunits may result in hydrophobic mismatch at the ER, reducing ER exit, and yield a conformation that is unfavorable for interaction with BK channel α subunits, thereby decreasing surface expression of BK channel α subunits (Chen et al., 2013).

Tools to analyze ion channel S-acylation

Before the seminal discovery of the mammalian enzymes that control S-acylation (Fukata et al., 2004) and current advances in proteomic techniques to assay S-acylation, progress in the field was relatively slow, largely because of the lack of pharmacological, proteomic, and genetic tools to investigate the functional role of S-acylation. It is perhaps instructive to consider that protein tyrosine phosphorylation was discovered the same year as S-acylation (Hunter, 2009). However, the subsequent rapid identification and cloning of tyrosine kinases provided a very extensive toolkit to investigate this pathway. Although the S-acylation toolkit remains limited, the last few years have seen rapid progress in our ability to interrogate S-acylation function and its control of ion channel physiology. Furthermore, S-acylation prediction algorithms, such as CSS-palm 4.0 (Ren et al., 2008), provide an in silico platform to inform experimental approaches for candidate targets.

Pharmacological tools.

The S-acylation pharmacological toolkit remains, unfortunately, empty, with limited specific agents with which to explore S-acylation function in vitro or in vivo. Although the palmitate analogue 2-bromopalmitate (2-BP) is widely used for cellular assays and to analyze ion channel regulation by S-acylation, caution must be taken in using this agent, even though it remains our best pharmacological inhibitor of zDHHCs (Resh, 2006; Davda et al., 2013; Zheng et al., 2013). Unfortunately, 2-BP is a nonselective inhibitor of lipid metabolism and many membrane-associated enzymes, and displays widespread promiscuity (e.g., Davda et al., 2013); does not show selectivity toward specific zDHHC proteins (Jennings et al., 2009); has many pleiotropic effects on cells at high concentrations, including cytotoxicity (Resh, 2006); and also inhibits acylthioesterases (Pedro et al., 2013). Other lipid inhibitors include cerulenin and tunicamycin. However, cerulenin affects many aspects of lipid metabolism, and tunicamycin inhibits N-linked glycosylation (Resh, 2006). Although some nonlipid inhibitors have been developed, these are not widely used (Ducker et al., 2006; Jennings et al., 2009), and there are currently no known activators of zDHHCs or compounds that inhibit specific zDHHCs. In the last few years, several inhibitors for the acylthioesterases LYPLA1 and LYPLA2 have been developed (Bachovchin et al., 2010; Dekker et al., 2010; Adibekian et al., 2012). However, several of these compounds, such as palmostatin B, are active against several members of the larger serine hydrolase family. Clearly, the development of novel S-acylation inhibitors and activators that display both specificity and zDHHC selectivity would represent a substantial advance for investigation of channel S-acylation.

Genetic tools.

To date, most studies have used overexpression of candidate zDHHCs in heterologous expression or native systems and analyzed increases in [3H]palmitate incorporation to define zDHHCs that may S-acylate specific ion channels (e.g. Rathenberg et al., 2004; Hayashi et al., 2005, 2009; Tian et al., 2010; Thomas et al., 2012). Although this is a powerful approach, caution is required to determine whether results obtained with overexpression in fact replicate endogenous regulation. For example, overexpression of some zDHHCs normally expressed in the cell type of interest can result in S-acylation of a cysteine residue that is not endogenously palmitoylated in BK channels (Tian et al., 2010). Point mutation of the cysteine of the catalytic DHHC domain abolishes the acyltransferase activity of zDHHCs and is thus an invaluable approach to confirming that the acyltransferase function of overexpressed zDHHC is required by itself. Increasingly, knockdown of endogenous zDHHCs using siRNA, and related approaches, is beginning to reveal the identity of zDHHCs that S-acylate native ion channel subunits. For example, knockdown of zDHHCs 5 or 8 reduces S-acylation of the accessory subunits PICK1 and Grip1, which control AMPA receptor trafficking (Thomas et al., 2012, 2013); and knockdown of zDHHC2 disrupts local nanoclusters of the PDZ domain protein PSD-95 in neuronal dendrites to control AMPA receptor membrane localization (Fukata et al., 2013). However, relatively few studies have taken a systematic knockdown approach to identify zDHHCs important for ion channel S-acylation. One such approach has, however, revealed that multiple, distinct zDHHCs mediate palmitoylation of the BK channel C terminus (zDHHCs 3, 5, 7, 9, and 17) and that a different subset of zDHHCs (22 and 23) mediate S-acylation of the intracellular S0-S1 loop in the same channel (Tian et al., 2010, 2012). Because some zDHHCs are themselves palmitoylated, the functional effect of overexpressing or knocking down individual zDHHCs on the localization and activity of other zDHHCs must also be carefully determined. For example, siRNA-mediated knockdown of zDHHC 5, 7, or 17 in HEK293 cells paradoxically results in an up-regulation of zDHHC23 mRNA expression (Tian et al., 2012). Furthermore, because many signaling and cytoskeletal elements are also controlled by S-acylation, direct effects on channel S-acylation by themselves must be evaluated in parallel (for example using site-directed cysteine mutants of the channel subunit). Fewer studies have used these approaches to examine the role of acylthioesterases, although overexpression of LYPLA1 and a splice variant of LYPLAL1, but not LYPLA2, deacylates the S0-S1 loop of the BK channel, promoting Golgi retention of the channels (Tian et al., 2012). Gene-trap and knockout mouse models for some zDHHCs (such as 5 and 17) are becoming available, although full phenotypic analysis and analysis of ion channel function in these models are largely lacking.

Proteomic and imaging tools. Lipid-centric (metabolic) labeling assays.

Metabolic labeling approaches are most suited to analysis of isolated cells, rather than tissues, but provide information on dynamic palmitoylation of proteins during the relatively short (∼4 h) labeling period as well as insight into the species of lipid bound to cysteine residues. The classical approach using radioactive palmitate (e.g., [3H]palmitate) remains a “gold standard” for validation, in particular for identification that palmitate is the bound lipid. However, metabolic labeling with [3H]palmitate generally requires immunoprecipitation and days to weeks of autoradiography or fluorography, particularly when analyzing low abundance membrane proteins such as ion channels. To overcome some of these issues, and also to provide a platform to allow cellular imaging of S-acylation, a variety of biorthogonal lipid probes have recently been developed (Hannoush and Arenas-Ramirez, 2009; Hannoush, 2012; Martin et al., 2012; for reviews see Charron et al., 2009a; Hannoush and Sun, 2010). These probes are modified fatty acids with reactive groups, such as an azide or alkyne group, allowing labeled proteins to be conjugated to biotin or fluorophores via the reactive group using Staudinger ligation or “click” chemistry. In particular, development of a family of ω-alkynyl fatty acid probes of different chain lengths (such as Alk-C16 and Alk-C18) have been exploited for proteomic profiling as well as single cell imaging (Gao and Hannoush, 2014) and have been used to identify candidate S-acylated channels in several mammalian cell lines (Charron et al., 2009b; Hannoush and Arenas-Ramirez, 2009; Martin and Cravatt, 2009; Yap et al., 2010; Yount et al., 2010; Martin et al., 2012). It is important to note that palmitic acid can also be incorporated into free N-terminal cysteines of proteins via an amide linkage (N-palmitoylation), addition of the monounsaturated palmitoleic acid via an oxyester linkage to a serine residue (O-palmitoylation), and oleic acid (oleoylation) as well as myristate via amide linkages on lysine residues (Stevenson et al., 1992; Linder and Deschenes, 2007; Hannoush and Sun, 2010; Schey et al., 2010). These modifications can be discriminated from S-acylation by their insensitivity to hydroxylamine cleavage (at neutral pH) compared with the S-acylation thioester linkage. Whether N- or O-linked palmitoylation or oleoylation controls ion channel function remains to be determined.

Cysteine centric (cysteine accessibility) assays: Acyl-biotin exchange (ABE) and resin-assisted capture (Acyl-RAC).

The metabolic labeling approach requires treating isolated cells with lipid conjugates and thus largely precludes analysis of native S-acylation in tissues. However, several related approaches have been developed that exploit the exposure of a reactive cysteine after hydroxylamine cleavage (at neutral pH) of the cysteine-acyl thioester linkage. The newly exposed cysteine thiol can then react with cysteine-reactive groups (such as biotin-BMCC or biotin-HPDP used in the ABE approach; Drisdel and Green, 2004; Drisdel et al., 2006; Draper and Smith, 2009; Wan et al., 2007) or thiopropyl sepharose (used in Acyl-RAC; Forrester et al., 2011) to allow purification of S-acylated proteins that can be identified by Western blot analysis or mass spectrometry. Acyl-RAC has been reported to improve detection of higher molecular weight S-acylated proteins and thus may prove valuable for ion channel analysis. These approaches have been exploited to determine the “palmitoylome” in several species and tissues (e.g., Wan et al., 2007, 2013; Kang et al., 2008; Martin and Cravatt, 2009; Yang et al., 2010; Yount et al., 2010; Merrick et al., 2011; Wilson et al., 2011; Jones et al., 2012; Ren et al., 2013). For example, analysis of rat brain homogenates identified both previously characterized as well as novel S-acylated ion channels (Wan et al., 2013), although it must be remembered that these approaches detect S-acylation and do not define S-palmitoylation per se. Cysteine accessibility approaches determine the net amount of preexisting S-acylated proteins; however, caution is required to eliminate false positives. In particular it is necessary to fully block all reactive cysteines before hydroxylamine cleavage; moreover, the identity of the endogenously bound lipid is of course not known.The lipid- and cysteine-centric approaches are thus complementary. In conjunction with site-directed mutagenesis of candidate S-acylated cysteine residues in ion channel subunits, these approaches have provided substantial insight into the role and regulation of ion channel S-acylation (Fukata et al., 2013). However, this approach does not directly confirm that the protein is S-acylated per se. Furthermore, in most ion channels, and in fact most S-acylated proteins, the identity of the native lipid bound to a specific S-acylated cysteine is not known. Although palmitate is considered to be the major lipid species involved in S-acylation, this has not been directly demonstrated in most cases, and other fatty acids, including arachidonic acid, oleate acid, and stearic acid, have also been reported to bind to cysteine via a thioester S-linkage (Linder and Deschenes, 2007; Hannoush and Sun, 2010). A major reason for this discrepancy is that mass spectrometry–based approaches to identify the native lipid specifically bound to S-acylated cysteines remain a significant challenge. This is particularly true for low abundance proteins such as mammalian ion channels, in contrast to the widespread application of mass spectrometry to directly identify native amino acids that are phosphorylated (Kordyukova et al., 2008, 2010; Sorek and Yalovsky, 2010; McClure et al., 2012; Ji et al., 2013). As such, direct biochemical demonstration of native cysteine S-acylation is lacking in most ion channels.

S-acylation and control of the ion channel lifecycle

Ion channel physiology is determined by both the number of channel proteins at the cognate membrane and by their activity and/or kinetics at the membrane. Evidence has begun to emerge that S-acylation of either pore-forming or regulatory subunits of ion channels controls all of these aspects of ion channel function. Although the focus of this review is S-acylation–dependent regulation of ion channel subunits itself, S-acylation also regulates the localization or activity of many adaptor, scaffolding, and cellular signaling proteins (e.g., G protein–coupled receptors [GPCRs], AKAP18, AKAP79/150, G proteins, etc.), as well as other aspects of cell biology that affect ion channel trafficking and the activity and regulation of macromolecular ion channel complexes (El-Husseini and Bredt, 2002; Linder and Deschenes, 2007; Fukata and Fukata, 2010; Greaves and Chamberlain, 2011; Shipston, 2011; Resh, 2012).

Control of ion channel cell surface expression and spatial organization in membranes.

The control of ion channel trafficking, from synthesis in the ER through modification in the Golgi apparatus to subsequent delivery to the appropriate cellular membrane compartment, is a major mechanism whereby S-acylation modulates ion channel physiology. S-acylation may influence the number of ion channels resident in a membrane through regulation of distinct steps in the ion channel lifecycle (Fig. 2). Indeed S-acylation has been implicated in ion channel synthesis, as well as in channel trafficking to the membrane and subsequent internalization, recycling, and degradation. S-acylation controls the maturation and correct assembly of ion channels early in the biosynthetic pathway. For example, S-acylation regulates assembly of the ligand gated nicotinic acetylcholine receptor (nAChR) to ensure a functional binding site for acetylcholine (Alexander et al., 2010) as well as controlling its surface expression (Amici et al., 2012). S-acylation is also an important determinant of the maturation of both voltage-gated sodium (Nav1.2) and voltage-gated potassium channels (Kv1.5; Schmidt and Catterall, 1987; Zhang et al., 2007). S-acylation also contributes to the efficient trafficking of channels from the ER to Golgi and to post-Golgi transport. Three examples illustrate the importance and potential complexity of S-acylation in controlling ion channel trafficking:(1) S-acylation of a cysteine residue adjacent to a hydrophobic region (TM2) in a cytosolic loop of the GluA1 pore-forming subunit of AMPA receptors (Fig. 4 A) promotes retention of the channel in the Golgi (Hayashi et al., 2005). However, S-acylated Grip1b, a PDZ protein that binds to AMPA receptors, is targeted to mobile trafficking vesicles in neuronal dendrites and accelerates local recycling of AMPA receptors to the plasma membrane (Thomas et al., 2012). In contrast, S-acylation of another AMPA receptor interacting protein, PICK1, is proposed to stabilize AMPA receptor internalization (Thomas et al., 2013).(2) S-acylation of a cluster of cysteine residues juxtaposed to the transmembrane 4 domain (Cys I site) of the NMDA receptor subunit GluN2A (Fig. 4 B) increases surface expression of NMDA receptors by decreasing their constitutive internalization. In contrast S-acylation at C-terminal cysteine residues (Cys II site) decreases their surface expression by introducing a Golgi retention signal that decreases forward trafficking (Hayashi et al., 2009). Even though both sites affect surface expression, only S-acylation of the TM4 juxtaposed cysteine residues influences synaptic incorporation of NMDA receptors, which suggests that this site is an important determinant of the synaptic versus extrasynaptic localization of these ion channels (Mattison et al., 2012). Together, these data highlight the importance of S-acylation of two distinct sites within the same ion channel as well as that of components of the ion channel multimolecular complex as determinants of channel trafficking.(3) S-acylation of a cluster of cysteine residues in the intracellular S0-S1 loop of the pore-forming subunit (Figs. 4 C and and5)5) is required for efficient exit of BK channels from the ER and the trans-Golgi network. Deacylation at the Golgi apparatus appears to be an important regulatory step (Tian et al., 2012). BK channel surface abundance may also be controlled by S-acylation of regulatory β4 subunits. β4 subunit S-acylation on a cysteine residue juxtaposed to the second transmembrane domain is important for the ability of the β4 subunit itself to exit the ER. Importantly, assembly of β4 subunits with specific splice variants of pore-forming α subunits of the BK channel enhances surface expression of the channel, a mechanism that depends on S-acylation of the β4 subunit (Chen et al., 2013). Thus, in BK channels, S-acylation of the S0-S1 loop of the pore-forming subunit controls global BK channel surface expression, and β4 subunit S-acylation controls surface expression of specific pore-forming subunit splice variants. S-acylation of the Kchip 2 and Kchip 3 accessory subunits also controls surface expression of voltage-gated Kv4.3 channels (Takimoto et al., 2002).Moreover, S-acylation modulates the spatial organization of ion channels within membranes. Perhaps the most striking example involves aquaporin 4 (AQP4), where S-acylation of two N-terminal cysteine residues in an N-terminal splice variant (AQP4M1) inhibits assembly of AQP4 into large orthogonal arrays (Suzuki et al., 2008; Crane and Verkman, 2009), perhaps by disrupting interactions within the AQP4 tetramer. S-acylation can affect the distribution of the many membrane-associated proteins between cholesterol-rich microdomains (lipid rafts) and the rest of the membrane. Such clustering has also been reported for various transmembrane proteins, including the P2x purinoceptor 7 (P2X7) receptor, in which S-acylation of the C terminus promotes clustering into lipid rafts (Gonnord et al., 2009). A similar mechanism may underlie synaptic clustering of GABAA receptors mediated by S-acylation of an intracellular loop of the y2 subunit (Rathenberg et al., 2004). In these examples, S-acylation of the channel itself affects membrane partitioning and organization. However, recent evidence in neurons suggests that establishment of “nano” domains of ion channel complexes in postsynaptic membranes may also be established by local clustering of the cognate acyltransferase itself. For example, clustering of zDHHC2 in the postsynaptic membranes of individual dendritic spines provides a mechanism for local control of S-acylation cycles of the PDZ protein adapter, PSD-95, and thereby for controlling its association with the plasma membrane. PSD-95, in turn, can assemble with various ion channels, including NMDA receptors, and can thus dynamically regulate the localization and clustering of ion channel complexes (Fukata et al., 2013). Indeed, an increasing number of other ion channel scaffolding proteins such as Grip1 (Thomas et al., 2012), PICK1 (Thomas et al., 2013), S-delphilin (Matsuda et al., 2006), and Ankyrin G (He et al., 2012) that influence ion channel trafficking, clustering, and localization are now known to be S-acylated.Relatively few studies have identified effects of S-acylation on the intrinsic gating kinetics or pharmacology of ion channels at the plasma membrane. However, a glycine-to-cysteine mutant (G1079C) in the intracellular loop between domains II and III enhances the sensitivity of the voltage-gated Na channel Nav1.2a to the toxins PaurTx3 and ProTx-II, an effect blocked by inhibition of S-acylation. These toxins control channel activation through the voltage sensor in domain III. In addition, deacylation of another (wild-type) cysteine residue (C1182) in the II–III loop produces a hyperpolarizing shift in both activation and steady-state inactivation as well as slowing the recovery from fast inactivation and increasing sensitivity to PaurTx3 (Bosmans et al., 2011). Effects of S-acylation on gating kinetics have also been reported in other channels. For example, in the voltage-sensitive potassium channel Kv1.1, S-acylation of the intracellular linker between transmembrane domains 2 and 3 increases the intrinsic voltage sensitivity of the channel (Gubitosi-Klug et al., 2005). S-acylation of the β and γ subunits of epithelial sodium channels (ENaC) also affects channel gating (Mueller et al., 2010; Mukherjee et al., 2014), and the S-acylated regulatory β2a subunit of N-type calcium channels controls voltage-dependent inactivation (Qin et al., 1998; Hurley et al., 2000).S-acylation is also an important determinant of retrieving ion channels from the plasma membrane for recycling or degradation. S-acylation of a single cysteine residue juxtaposed to the transmembrane TM4 domain of GluA1 and GluA2 subunits of AMPA receptors controls agonist-induced ion channel internalization. These residues are distinct from those controlling Golgi retention of AMPA receptors (Fig. 4 A), which emphasizes the finding that the location and context of the S-acylated cysteines, even in the same protein, is central for their effects on physiological function (Hayashi et al., 2005; Lin et al., 2009; Yang et al., 2009). The stability of many proteins is also regulated by S-acylation; S-acylation of a single cysteine residue in Kv1.5 promotes both its internalization and its degradation (Zhang et al., 2007; Jindal et al., 2008). Thus, in different ion channels, S-acylation can have opposite effects on insertion, membrane stability, and retrieval.

S-acylation and posttranslational cross-talk control channel trafficking and activity.

An emerging concept is that S-acylation is an important determinant of ion channel regulation by other PTMs. Indeed, nearly 20 years ago it was reported that PKC-dependent phosphorylation of the GluK2 (GluR6) subunit of Kainate receptors was attenuated in channels S-acylated at cysteine residues near the PKC consensus site (Pickering et al., 1995). S-acylation of GluA1 subunits of AMPA receptors also blocks PKC phosphorylation of GluA1 and subsequently prevents its binding to the cytoskeletal adapter protein 4.1N, ultimately disrupting AMPA receptor insertion into the plasma membrane (Lin et al., 2009). Intriguingly, PKC phosphorylation and S-acylation have the opposite effect on 4.1N-mediated regulation of Kainate receptor (GluK2 subunit) membrane insertion: in this, case S-acylation promotes 4.1N interaction with Kainate receptors and thereby receptor insertion, whereas PKC phosphorylation disrupts 4.1N interaction, promoting receptor internalization (Copits and Swanson, 2013). Disruption of phosphorylation by S-acylation of residues near consensus phosphorylation sites likely results from steric hindrance, as proposed for S-acylation–dependent regulation of β2 adrenergic receptor phosphorylation (Mouillac et al., 1992; Moffett et al., 1993).S-acylation has also been reported to promote ion channel phosphorylation. For example, site-directed mutation of a cluster of palmitoylated cysteine residues in the GluN2A subunit of NMDA receptors abrogates Fyn-dependent tyrosine phosphorylation at a site between TM4 and the palmitoylated cysteines (Hayashi et al., 2009). Therefore, S-acylation of GluN2A promotes tyrosine phosphorylation, resulting in reduced internalization of the NMDA receptor (Hayashi et al., 2009). Furthermore, S-acylation of BK channels can act as a gate to switch channel regulation to different AGC family kinase signaling pathways, emphasizing the complex interactions that can occur between signaling pathways (Tian et al., 2008; Zhou et al., 2012; Fig. 5). S-acylation of an alternatively spliced insert (STREX) in the large cytosolic domain of the pore-forming subunit of BK channels promotes association of the STREX domain with the plasma membrane. S-acylation of the STREX insert is essential for the functional inhibition of STREX BK channels by PKA-mediated phosphorylation of a serine residue immediately upstream of the S-acylated cysteines. PKA phosphorylation dissociates the STREX domain from the plasma membrane (Tian et al., 2008), preventing STREX domain S-acylation (Jeffries et al., 2012) and leading to channel inhibition. However, deacylation of the STREX domain exposes a PKC consensus phosphorylation site downstream of the STREX domain, allowing PKC to inhibit STREX BK channels (Zhou et al., 2012). Thus, S-acylation acts as a reversible switch to specify regulation by AGC family kinases through control of the membrane association of a cytosolic domain of the channel: S-acylated STREX BK channels are inhibited by PKA but insensitive to PKC, whereas deacylated channels are inhibited by PKC but not PKA (Fig. 5). The reciprocal control of membrane association of a protein domain by S-acylation and protein phosphorylation likely represents a common mechanism in other signaling proteins as revealed for phosphodiesterase 10A (Charych et al., 2010).Cysteine residues are targets for several other modifications that regulate various ion channels, including nitrosylation, sulphydration, REDOX regulation, and formation of disulphide bonds (Sen and Snyder, 2010). Evidence is beginning to emerge that S-acylation may mutually compete with these mechanisms, providing a dynamic network to control cysteine reactivity. For example, the ion channel scaffolding PDZ domain protein PSD-95 is S-acylated at two N-terminal cysteine residues (C3 and C5) that are required for membrane targeting and clustering of PSD-95 (El-Husseini et al., 2002). nNOS also interacts with PSD-95, and stimulation of nitric oxide production results in nitrosylation of these cysteines, preventing their S-acylation and thereby decreasing PSD-95 clusters at postsynaptic sites (Ho et al., 2011). A recent remarkable example of the potential for such cross-talk in ion channel subunits is the identification of the S-acylation of 18 different cysteine residues in the large cytosolic N terminus of RyR1 in skeletal muscle. Of these 18 S-acylated cysteines, six have previously been identified as targets for S-oxidation, and a further cysteine residue was also subject to S-nitrosylation (Chaube et al., 2014) Although the functional relevance of this potential cross-talk in RyR1 has yet to be defined, interaction between oxidation and S-acylation of the same cysteine residue is physiologically relevant in other proteins. For example, oxidation of the signaling protein HRas at two cysteine residues C181/184 prevents S-acylation of these residues, resulting in a loss of plasma membrane localization of this peripheral membrane signaling protein (Burgoyne et al., 2012). Intriguingly, a conserved cysteine residue in nAChR α3 subunits, which has been shown to be S-acylated (C273) in the nAChR α4 subunit, has been implicated in use-dependent inactivation of nAChRs by reactive oxygen species (Amici et al., 2012). Determining whether these mutually competitive cysteine modifications represent an important mechanism for regulation of a range of ion channels is an exciting challenge for the future.S-acylation is also an important determinant of ion channel regulation by heterotrimeric G proteins. This can involve S-acylation of either G protein targets or of regulators of G proteins. In an example of the former, the palmitoylated N terminus of the regulatory β2a subunit splice variant acts as a steric inhibitor of an arachidonic acid binding domain to stimulate N-type calcium channels (Chien et al., 1996; Heneghan et al., 2009; Mitra-Ganguli et al., 2009). When the regulatory β subunits are not S-acylated, however, Gq-mediated signaling, via arachidonic acid, inhibits calcium channel activity. Closure of G protein regulated inward rectifying potassium (GIRK) channels in neurons after Gi/o deactivation provides an example of the latter (Jia et al., 2014). Signaling by members of the Gi/o family of the Gα subunit of heterotrimeric G proteins is terminated by members of the regulator of G protein signaling 7 (R7 RGS) family of GTPase-activating proteins, which accelerate GTP hydrolysis to speed Gi/o deactivation. Membrane localization of regulator of G protein signaling 7 (R7-RGS) is required for its regulation of Gi/o, and this is determined by interaction with an S-acylated R7 binding protein (R7-BP) that acts as an allosteric activator. Thus, the R7-RGS complex, recruited to the plasma membrane by S-acylated R7-BP, promotes Gi/o deactivation to facilitate GIRK channel closure. Conversely, deacylation of R7-BP removes the R7-GS complex from the plasma membrane, slowing Gi/o deactivation and consequent channel closure (Jia et al., 2014). Clearly, as S-acylation can also control an array of GPCRs, enzymes, and signaling and adapter proteins that indirectly control ion channel function (El-Husseini and Bredt, 2002; Linder and Deschenes, 2007; Fukata and Fukata, 2010; Greaves and Chamberlain, 2011; Shipston, 2011; Resh, 2012), understanding how S-acylation dynamically controls other components of ion channel multimolecular signaling complexes will be an essential future goal.

Summary and perspectives

With an ever-expanding “catalog” of S-acylated ion channel pore-forming and regulatory subunits (∼50 to date), together with an array of S-acylated scaffolding and signaling proteins, the importance and ubiquity of this reversible covalent lipid modification in controlling the lifecycle and physiological function and regulation of ion channels is unquestionable. This has been paralleled by a major resurgence in the wider S-acylation field, a consequence in large part of the discovery of S-acylating and deacylating enzymes together with a growing arsenal of genetic, proteomic, imaging, and pharmacological tools to assay and interrogate S-acylation function.As for most other posttranslational modifications of ion channels, including phosphorylation, major future goals for the field include:(1) Understanding mechanistically how covalent addition of a fatty acid can control such a diverse array of ion channel protein properties and functions, and how this is spatiotemporally regulated.(2) Elucidating the physiological relevance of this posttranslational modification from the level of single ion channels to the functional role of the channel in the whole organism in health and disease.Elucidation of these issues has fundamental implications far beyond ion channel physiology.To address these goals several major challenges and questions must be addressed, including:(1) It is largely assumed that S-acylation of transmembrane proteins results in an additional “membrane anchor” to target domains to the membrane interface. However, understanding the mechanisms, forces, and impact of S-acylation on the orientation of transmembrane helices and the architecture and structure of disordered domains in cytosolic loops and linkers, while remaining a considerable technical challenge, should provide major insight into mechanisms controlling channel trafficking, activity, and regulation.(2) Although S-acylation is widely accepted to be reversible, its spatiotemporal regulation of most ion channels is unknown. Mechanistic insight into zDHHC and acylthioesterase substrate specificity, native subcellular localization, and assembly with ion channel signaling complexes will allow us to dissect and understand how S-acylation of ion channels is controlled. Importantly, this should allow us to take both “channel-centric” (e.g., site-directed mutagenesis of S-acylated cysteines) as well as “S-acylation centric” (e.g., knockout of specific zDHHC activity) approaches to understand how multisite S-acylation on the same ion channel subunit can control distinct functions as well as physiological regulation of trafficking and function at the plasma membrane.(3) The functional role of S-acylation cannot be viewed in isolation from other posttranslational modifications. The cross-talk between S-acylation and adjacent phosphorylation sites as well as other cysteine modifications highlights the importance of understanding the interactions between signaling pathways. Insight into the rules, mechanisms, and cross-talk of S-acylation with these modifications has broad implications for cellular signaling.(4) Although it is clear that disruption of S-acylation homeostasis itself has substantial effects on normal physiology, and we are beginning to understand some of the cellular functions of ion channel S-acylation, we know very little about the functional impact of disrupted ion channel S-acylation at the systems and organismal level. Understanding how this may be dynamically regulated during a lifespan is critical to understanding the role of S-acylation in health and disease.To address these issues, development of improved tools to assay and investigate S-acylation from the single protein to organism is required. For example, tools to allow the real-time analysis of S-acylation status of ion channels in cells and tissues will provide fundamental insights into its dynamics and role in ion channel trafficking and membrane localization. Improved proteomic tools will allow direct assay of fatty acids bound to cysteine residues via thioester linkages. Development of new tools and models are essential if we are to understand the physiological relevance of ionic channel S-acylation at the systems level. These include: specific inhibitors of zDHHCs and thioesterases, conditional knockouts to spatiotemporally control zDHHC expression, and transgenics expressing catalytically inactive zDHHCs and models expressing S-acylation–null ion channel subunits. Furthermore, our understanding of how S-acylation may be dynamically controlled during normal ageing in response to homeostatic challenge and disruption in disease states remains rudimentary. Whether we will start to uncover channel “S-acylationopathies” resulting from dysregulation of ion channel S-acylation, analogous to channel phosphorylopathies, remains to be explored. Addressing these issues, together with development of new tools, will provide a paradigm shift in our understanding of both ion channel and S-acylation physiology, and promises to reveal novel therapeutic strategies for a diverse array of disorders.  相似文献   

13.
Evolution and Function of the Plant Cell Wall Synthesis-Related Glycosyltransferase Family 8     
Yanbin Yin  Huiling Chen  Michael G. Hahn  Debra Mohnen  Ying Xu 《Plant physiology》2010,153(4):1729-1746
Carbohydrate-active enzyme glycosyltransferase family 8 (GT8) includes the plant galacturonosyltransferase1-related gene family of proven and putative α-galacturonosyltransferase (GAUT) and GAUT-like (GATL) genes. We computationally identified and investigated this family in 15 fully sequenced plant and green algal genomes and in the National Center for Biotechnology Information nonredundant protein database to determine the phylogenetic relatedness of the GAUTs and GATLs to other GT8 family members. The GT8 proteins fall into three well-delineated major classes. In addition to GAUTs and GATLs, known or predicted to be involved in plant cell wall biosynthesis, class I also includes a lower plant-specific GAUT and GATL-related (GATR) subfamily, two metazoan subfamilies, and proteins from other eukaryotes and cyanobacteria. Class II includes galactinol synthases and plant glycogenin-like starch initiation proteins that are not known to be directly involved in cell wall synthesis, as well as proteins from fungi, metazoans, viruses, and bacteria. Class III consists almost entirely of bacterial proteins that are lipooligo/polysaccharide α-galactosyltransferases and α-glucosyltransferases. Sequence motifs conserved across all GT8 subfamilies and those specific to plant cell wall-related GT8 subfamilies were identified and mapped onto a predicted GAUT1 protein structure. The tertiary structure prediction identified sequence motifs likely to represent key amino acids involved in catalysis, substrate binding, protein-protein interactions, and structural elements required for GAUT1 function. The results show that the GAUTs, GATLs, and GATRs have a different evolutionary origin than other plant GT8 genes, were likely acquired from an ancient cyanobacterium (Synechococcus) progenitor, and separate into unique subclades that may indicate functional specialization.Plant cell walls are composed of three principal types of polysaccharides: cellulose, hemicellulose, and pectin. Studying the biosynthesis and degradation of these biopolymers is important because cell walls have multiple roles in plants, including providing structural support to cells and defense against pathogens, serving as cell-specific developmental and differentiation markers, and mediating or facilitating cell-cell communication. In addition to their important roles within plants, cell walls also have many economic uses in human and animal nutrition and as sources of natural textile fibers, paper and wood products, and components of fine chemicals and medicinal products. The study of the biosynthesis and biodegradation of plant cell walls has become even more significant because cell walls are the major components of biomass (Mohnen et al., 2008), which is the most promising renewable source for the production of biofuels and biomaterials (Ragauskas et al., 2006; Pauly and Keegstra, 2008). Analyses of fully sequenced plant genomes have revealed that they encode hundreds or even thousands of carbohydrate-active enzymes (CAZy; Henrissat et al., 2001; Yokoyama and Nishitani, 2004; Geisler-Lee et al., 2006). Most of these CAZy enzymes (Cantarel et al., 2009) are glycosyltransferases (GTs) or glycoside hydrolases, which are key players in plant cell wall biosynthesis and modification (Cosgrove, 2005).The CAZy database is classified into 290 protein families (www.cazy.org; release of September 2008), of which 92 are GT families (Cantarel et al., 2009). A number of the GT families have been previously characterized to be involved in plant cell wall biosynthesis. For example, the GT2 family is known to include cellulose synthases and some hemicellulose backbone synthases (Lerouxel et al., 2006), such as mannan synthases (Dhugga et al., 2004; Liepman et al., 2005), putative xyloglucan synthases (Cocuron et al., 2007), and mixed linkage glucan synthases (Burton et al., 2006). With respect to the synthesis of xylan, a type of hemicellulose, four Arabidopsis (Arabidopsis thaliana) proteins from the GT43 family, irregular xylem 9 (IRX9), IRX14, IRX9-L, and IRX14-L, and two proteins from the GT47 family, IRX10 and IRX10-L, are candidates (York and O''Neill, 2008) for glucuronoxylan backbone synthases (Brown et al., 2007, 2009; Lee et al., 2007a; Peña et al., 2007; Wu et al., 2009). In addition, three proteins have been implicated in the synthesis of an oligosaccharide thought to act either as a primer or terminator in xylan synthesis (Peña et al., 2007): two from the GT8 family (IRX8/GAUT12 [Persson et al., 2007] and PARVUS/GATL1 [Brown et al., 2007; Lee et al., 2007b]) and one from the GT47 family (FRA8/IRX7 [Zhong et al., 2005]).The GT families involved in the biosynthesis of pectins have been relatively less studied until recently. In 2006, a gene in CAZy family GT8 was shown to encode a functional homogalacturonan α-galacturonosyltransferase, GAUT1 (Sterling et al., 2006). GAUT1 belongs to a 25-member gene family in Arabidopsis, the GAUT1-related gene family, that includes two distinct but closely related families, the galacturonosyltransferase (GAUT) genes and the galacturonosyltransferase-like (GATL) genes (Sterling et al., 2006). Another GAUT gene, GAUT8/QUA1, has been suggested to be involved in pectin and/or xylan synthesis, based on the phenotypes of plant lines carrying mutations in this gene (Bouton et al., 2002; Orfila et al., 2005). It has further been suggested that multiple members of the GT8 family are galacturonosyltransferases involved in pectin and/or xylan biosynthesis (Mohnen, 2008; Caffall and Mohnen, 2009; Caffall et al., 2009).Aside from the 25 GAUT and GATL genes, Arabidopsis has 16 other family GT8 genes, according to the CAZy database, which do not seem to have the conserved sequence motifs found in GAUTs and GATLs: HxxGxxKPW and GLG (Sterling et al., 2006). Eight of these 16 genes are annotated as galactinol synthase (GolS) by The Arabidopsis Information Resource (TAIR; www.arabidopsis.org), and three of these AtGolS enzymes have been implicated in the synthesis of raffinose family oligosaccharides that are associated with stress tolerance (Taji et al., 2002). The other eight Arabidopsis GT8 genes are annotated as plant glycogenin-like starch initiation proteins (PGSIPs) in TAIR. PGSIPs have been proposed to be involved in the synthesis of primers necessary for starch biosynthesis (Chatterjee et al., 2005). Hence, the GT8 family is a protein family consisting of enzymes with very distinct proven and proposed functions. Indeed, a suggestion has been made to split the GT8 family into two groups (Sterling et al., 2006), namely, the cell wall biosynthesis-related genes (GAUTs and GATLs) and the non-cell wall synthesis-related genes (GolSs and PGSIPs).We are interested in further defining the functions of the GAUT and GATL proteins in plants, in particular their role(s) in plant cell wall synthesis. The apparent disparate functions of the GT8 family (i.e. the GAUTs and GATLs as proven and putative plant cell wall polysaccharide biosynthetic α-galacturonosyltransferases, the eukaryotic GolSs as α-galactosyltransferases that synthesize the first step in the synthesis of the oligosaccharides stachyose and raffinose, the putative PGSIPs, and the large bacterial GT8 family of diverse α-glucosyltransferases and α-galactosyltransferases involved in lipopolysaccharide and lipooligosaccharide synthesis) indicate that the GT8 family members are involved in several unique types of glycoconjugate and glycan biosynthetic processes (Yin et al., 2010). This observation led us to ask whether any of the GT8 family members are sufficiently closely related to GAUT and GATL genes to be informative regarding GAUT or GATL biosynthetic function(s) and/or mechanism(s).To investigate the relatedness of the members of the GT8 gene family, we carried out a detailed phylogenetic analysis of the entire GT8 family in 15 completely sequenced plant and green algal genomes (AbbreviationCladeSpeciesGenome PublishedDownloaded frommpcGreen algaeMicromonas pusilla CCMP1545Worden et al. (2009)JGI version 2.0mprGreen algaeMicromonas strain RCC299Worden et al. (2009)JGI version 2.0olGreen algaeOstreococcus lucimarinusPalenik et al. (2007)JGI version 1.0otGreen algaeOstreococcus tauriDerelle et al. (2006)JGI version 1.0crGreen algaeChlamydomonas reinhardtiiMerchant et al. (2007)JGI version 3.0vcGreen algaeVolvox carteri f. nagariensisNoJGI version 1.0ppMossPhyscomitrella patens ssp. patensRensing et al. (2008)JGI version 1.1smSpike mossSelaginella moellendorffiiNoJGI version 1.0ptDicotPopulus trichocarpaTuskan et al. (2006)JGI version 1.1atDicotArabidopsis thalianaArabidopsis Genome Initiative (2000)TAIR version 9.0vvDicotVitis viniferaJaillon et al. (2007)http://www.genoscope.cns.fr/gmDicotGlycine maxSchmutz et al. (2010)JGI version 1.0osMonocotOryza sativaGoff et al. (2002); Yu et al. (2002)TIGR version 6.1sbMonocotSorghum bicolorPaterson et al. (2009)JGI version 1.0bdMonocotBrachypodium distachyonVogel et al. (2010)JGI version 1.0Open in a separate window  相似文献   

14.
On the Classification of Epistatic Interactions     
Hong Gao  Julie M. Granka  Marcus W. Feldman 《Genetics》2010,184(3):827-837
Modern genomewide association studies are characterized by the problem of “missing heritability.” Epistasis, or genetic interaction, has been suggested as a possible explanation for the relatively small contribution of single significant associations to the fraction of variance explained. Of particular concern to investigators of genetic interactions is how to best represent and define epistasis. Previous studies have found that the use of different quantitative definitions for genetic interaction can lead to different conclusions when constructing genetic interaction networks and when addressing evolutionary questions. We suggest that instead, multiple representations of epistasis, or epistatic “subtypes,” may be valid within a given system. Selecting among these epistatic subtypes may provide additional insight into the biological and functional relationships among pairs of genes. In this study, we propose maximum-likelihood and model selection methods in a hypothesis-testing framework to choose epistatic subtypes that best represent functional relationships for pairs of genes on the basis of fitness data from both single and double mutants in haploid systems. We gauge the performance of our method with extensive simulations under various interaction scenarios. Our approach performs reasonably well in detecting the most likely epistatic subtype for pairs of genes, as well as in reducing bias when estimating the epistatic parameter (ɛ). We apply our approach to two available data sets from yeast (Saccharomyces cerevisiae) and demonstrate through overlap of our identified epistatic pairs with experimentally verified interactions and functional links that our results are likely of biological significance in understanding interaction mechanisms. We anticipate that our method will improve detection of epistatic interactions and will help to unravel the mysteries of complex biological systems.UNDERSTANDING the nature of genetic interactions is crucial to obtaining a more complete picture of complex biological systems and their evolution. The discovery of genetic interactions has been the goal of many researchers studying a number of model systems, including but not limited to Saccharomyces cerevisiae, Caenorhabditis elegans, and Escherichia coli (You and Yin 2002; Burch et al. 2003; Burch and Chao 2004; Tong et al. 2004; Drees et al. 2005; Sanjuán et al. 2005; Segre et al. 2005; Pan et al. 2006; Zhong and Sternberg 2006; Jasnos and Korona 2007; St. Onge et al. 2007; Decourty et al. 2008). Recently, high-throughput experimental approaches, such as epistatic mini-array profiles (E-MAPs) and genetic interaction analysis technology for E. coli (GIANT-coli), have enabled the study of epistasis on a large scale (Schuldiner et al. 2005, 2006; Collins et al. 2006, 2007; Typas et al. 2008). However, it remains unclear whether the computational and statistical methods currently in use to identify these interactions are indeed the most appropriate.The study of genetic interaction, or “epistasis,” has had a long and somewhat convoluted history. Bateson (1909) first used the term epistasis to describe the ability of a gene at one locus to “mask” the mutational influence of a gene at another locus (Cordell 2002). The term “epistacy” was later coined by Fisher (1918) to denote the statistical deviation of multilocus genotype values from an additive linear model for the value of a phenotype (Phillips 1998, 2008).These origins are the basis for the two main current interpretations of epistasis. The first, as introduced by Bateson (1909), is the “biological,” “physiological,” or “compositional” form of epistasis, concerned with the influence of an individual''s genetic background on an allele''s effect on phenotype (Cheverud and Routman 1995; Phillips 1998, 2008; Cordell 2002; Moore and Williams 2005). The second interpretation, attributed to Fisher, is “statistical” epistasis, which in its linear regression framework places the phenomenon of epistasis in the context of a population (Wagner et al. 1998; Wade et al. 2001; Wilke and Adami 2001; Moore and Williams 2005; Phillips 2008). Each of these approaches is equally valid in studying genetic interactions; however, confusion still exists about how to best reconcile the methods and results of the two (Phillips 1998, 2008; Cordell 2002; Moore and Williams 2005; Liberman and Feldman 2006; Aylor and Zeng 2008).Aside from the distinction between the statistical and the physiological definitions of epistasis, inconsistencies exist when studying solely physiological epistasis. For categorical traits, physiological epistasis is clear as a “masking” effect. When noncategorical or numerical traits are measured, epistasis is defined as the deviation of the phenotype of the multiple mutant from that expected under independence of the underlying genes.The “expectation” of the phenotype under independence, that is, in the absence of epistasis, is not defined consistently between studies. For clarity, consider epistasis between pairs of genes and, without loss of generality, consider fitness as the phenotype. The first commonly used definition of independence, originating from additivity, defines the effect of two independent mutations to be equal to the sum of the individual mutational effects. A second, motivated by the use of fitness as a phenotype, defines the effect of the two mutations as the product of the individual effects (Elena and Lenski 1997; Desai et al. 2007; Phillips 2008). A third definition of independence has been referred to as “minimum,” where alleles at two loci are independent if the double mutant has the same fitness as the less-fit single mutant. Mani et al. (2008) claim that this has been used when identifying pairwise epistasis by searching for synthetic lethal double mutants (Tong et al. 2001, 2004; Pan et al. 2004, 2006; Davierwala et al. 2005). A fourth is the “Log” definition presented by Mani et al. (2008) and Sanjuan and Elena (2006). The less-frequently used “scaled ɛ” (Segre et al. 2005) measure of epistasis takes the multiplicative definition of independence with a scaling factor.These different definitions of independence are partly due to distinct measurement “scales.” For some traits, a multiplicative definition of independence may be necessary to identify epistasis between two genes, whereas for other traits, additivity may be appropriate (Falconer and Mackay 1995; Wade et al. 2001; Mani et al. 2008; Phillips 2008). An interaction found under one independence definition may not necessarily be found under another, leading to different biological conclusions (Mani et al. 2008).Mani et al. (2008) suggest that there may be an “ideal” definition of independence for all gene pairs for identifying functional relationships. However, it is plausible that different representations of independence for two genes may reflect different biological properties of the relationship (Kupper and Hogan 1978; Rothman et al. 1980). “Two categories of general interest [the additive and multiplicative definitions, respectively] are those in which etiologic factors act interchangeably in the same step in a multistep process, or alternatively act at different steps in the process” (Rothman et al. 1980, p. 468). In some cases, the discovery of epistasis may merely be an artifact of using an incorrect null model (Kupper and Hogan 1978). It may be necessary to represent “independence” differently, resulting in different statistical measures of interactions, for different pairs of genes depending on their functions.Previous studies have suggested that different pairs of loci may have different modes of interaction and have attempted to subclassify genetic interactions into regulatory hierarchies and mutually exclusive “interaction subtypes” to elucidate underlying biological properties (Avery and Wasserman 1992; Drees et al. 2005; St. Onge et al. 2007). We suggest that epistatic relationships can be divided into several subtypes, or forms, corresponding to the aforementioned definitions of independence. As a particular gene pair may deviate from independence according to several criteria, we do not claim that these subtypes are necessarily mutually exclusive. We attempt to select the most likely epistatic subtype that is the best statistical representation of the relationship between two genes. To further subclassify interactions, epistasis among deleterious mutations can take one of two commonly used forms: positive (equivalently alleviating, antagonistic, or buffering) epistasis, where the phenotype of the double mutant is less severe than expected under independence, and negative (equivalently aggravating, synergistic, or synthetic), where the phenotype is more severe than expected (Segre et al. 2005; Collins et al. 2006; Desai et al. 2007; Mani et al. 2008).Another objective of such distinctions is to reduce the bias of the estimator of the epistatic parameter (ɛ), which measures the extent and direction of epistasis for a given gene pair. Mani et al. (2008), assuming that the overall distribution of ɛ should be centered around 0, find that inaccurately choosing a definition of independence can result in increased bias when estimating ɛ. For example, using the minimum definition results in the most severe bias when single mutants have moderate fitness effects, and the additive definition results in the largest positive bias when at least one gene has an extreme fitness defect (Mani et al. 2008). Therefore, it is important to select an optimal estimator for ɛ for each pair of genes from among the subtypes of epistatic interactions.Epistasis may be important to consider in genomic association studies, as a gene with a weak main effect may be identified only through its interaction with another gene or other genes (Frankel and Schork 1996; Culverhouse et al. 2002; Moore 2003; Cordell 2009; Moore and Williams 2009). Epistasis has also been studied extensively in the context of the evolution of sex and recombination. The mutational deterministic hypothesis proposes that the evolution of sex and recombination would be favored by negative epistatic interactions (Feldman et al. 1980; Kondrashov 1994); many other studies have also studied the importance of the form of epistasis (Elena and Lenski 1997; Otto and Feldman 1997; Burch and Chao 2004; Keightley and Otto 2006; Desai et al. 2007; MacCarthy and Bergman 2007). Indeed, according to Mani et al. (2008, p. 3466), “the choice of definition [of epistasis] alters conclusions relevant to the adaptive value of sex and recombination.”Given fitness data from single and double mutants in haploid organisms, we implement a likelihood method to determine the subtype that is the best statistical representation of the epistatic interaction for pairs of genes. We use maximum-likelihood estimation and the Bayesian information criteria (BIC) (Schwarz 1978) with a likelihood-ratio test to select the most appropriate null or epistatic model for each putative interaction. We conduct extensive simulations to gauge the performance of our method and demonstrate that it performs reasonably well under various interaction scenarios. We apply our method to two data sets with fitness measurements obtained from yeast (Jasnos and Korona 2007; St. Onge et al. 2007), whose authors assume only multiplicative epistasis for all interactions. By examining functional links and experimentally validated interactions among epistatic pairs, we demonstrate that our results are biologically meaningful. Studying a random selection of genes, we find that minimum epistasis is more prevalent than both additive and multiplicative epistasis and that the overall distribution of ɛ is not significantly different from zero (as Jasnos and Korona 2007 suggest). For genes in a particular pathway, we advise selecting among fewer epistatic subtypes. We believe that our method of epistatic subtype classification will aid in understanding genetic interactions and their properties.

St. Onge et al. (2007) data set:

St. Onge et al. (2007) examined 26 nonessential genes known to confer resistance to MMS, constructed double-deletion strains for 323 double-mutant strains (all but two of the total possible pairs), and assumed the multiplicative form of epistasis for all interactions (see Methods: Analysis of experimental data). Following these authors, we focus on single- and double-mutant fitnesses measured in the presence of MMS. (For results in the absence of MMS, see File S1 and File S1_2.)Using the resampling method described in Analysis of experimental data and File S1, 222 gene pairs pass the cutoff of having epistasis inferred in at least 900 of 1000 replicates. This does not include 5 synthetic lethal gene pairs. Hypothesis testing and a multiple-testing procedure (for 222 simultaneous hypotheses) are necessary to determine the final epistatic pairs.To select one among the three multiple-testing procedures, we follow St. Onge et al. (2007) and examine gene pairs that share specific functional links (see Analysis of experimental data). The Bonferroni method is likely too conservative, yielding only 25 significantly epistatic pairs with only one functional link among them; alternatively, the pFDR procedure appears to be too lenient in rejecting independence for all 222 pairs. Therefore, we use the FDR procedure (although the number of functional links is not significant) and detect 193 epistatic pairs, of which 5 (2.6%) are synthetic lethals, 19 (9.8%) have additive epistasis, 33 (17.1%) have multiplicative epistasis, and 136 (70.5%) have minimum epistasis (File S1_1). We find 29 gene pairs with positive (alleviating) epistasis and 159 gene pairs with negative (aggravating) epistasis.

TABLE 2

Summary of gene pairs with the indicated epistatic subtypes, inferred using the FDR procedure with the BIC method that considers all three epistatic subtypes and their corresponding null models
Epistatic subtypeStudy SStudy J
All193 (100%)352 (100%)
= −0.060 = −0.001
= −0.096 = −0.059
Additive19 (9.8%)35 (9.9%)
= 0.115* = 0.193***
= 0.131 = 0.188
Multiplicative33 (17.1%)63 (17.9%)
= 0.048 = 0.017
= −0.166 = −0.115
Minimum136 (70.5%)254 (72.2%)
= −0.111*** = −0.032**
= −0.091 = −0.065
Open in a separate windowNumbers are the counts of each type, and percentages are given of the total number of epistatic pairs. The mean () and median () of the epistatic parameter (ɛ) are given for each subtype, with “*” indicating that the mean of ɛ is significantly different from 0 (*, P-value ≤0.05; **, P-value ≤0.01; ***, P-value ≤0.001). Study S refers to the St. Onge et al. (2007) data set, and study J refers to the Jasnos and Korona (2007) data set. (For study S, five of the epistatic pairs are synthetic lethals and are not shown; as a result, percentages do not sum to 100%.)To further validate the use of our method and the FDR procedure, we assess by Fisher''s exact test the significance of an enrichment of both Biological Process and all GO Slim term links among epistatic pairs, neither of which are significant (Gene Ontology Consortium 2000; www.yeastgenome.org; Stark et al. 2006); Table S4]. Although some of the previously unidentified interactions that we identify could be false positives, many are likely to be new discoveries.

TABLE 3

Comparison of validation measures for each data set for different variations of the FDR and BIC procedures, considering only a subset of epistatic subtypes with their corresponding null models: all epistatic subtypes (A, P, and M); only the additive and multiplicative subtypes (A and P); and only the additive (A), only the multiplicative (P), or only the minimum (M) subtype (see text for details)
Subtypes considered in BIC procedure
A, P, MA, PAPM
Study J
No. found (636)352273263231329
Functional links (25)19 (0.0255)*13 (0.2320)11 (0.4689)10 (0.4227)15 (0.2619)
GO Slim terms (Biological Process) (115)69 (0.1573)50 (0.4874)55 (0.0736)44 (0.3534)68 (0.04902)*
GO Slim terms (all) (369)224 (0.0009)*172 (0.01654)*160 (0.1297)146 (0.0273)*213 (0.0003)*
Experimentally identified (3)32123
Study S
No. found (323)193192247171243
Functional links (36)21 (0.6450)29 (0.0041)*34 (0.0031)*29 (0.0003)*24 (0.9256)
GO Slim terms (Biological Process) (283)174 (0.0657)174 (0.03656)*223 (0.0010)*153 (0.1825)213 (0.5534)
GO Slim terms (all) (307)185 (0.2866)182 (0.6926)237 (0.1472)162 (0.6997)231 (0.5908)
Experimentally identified (29)1722242321
Open in a separate windowNumbers in parentheses indicate P-values by Fisher''s exact test. “*” indicates significance. Study J refers to the Jasnos and Korona (2007) data set, and study S refers to the St. Onge et al. (2007) data set measured in the presence of MMS. Numbers in parentheses indicate the total number of tested pairs and the total number of each type of link found in each complete data set.The epistatic subtypes we consider are not necessarily mutually exclusive. To more fully assess the assumptions of our method, we also consider several of the possible subsets of the epistatic subtypes (and their corresponding null models) in our procedure. As the minimum epistatic subtype was the most frequently selected in this data set, we first do not include the minimum null model or the minimum epistatic model in our procedure (i.e., we select from among four rather than six models for a pair; Table S4). However, there are a significant number of epistatic pairs with functional links only when the minimum epistatic subtype is not included (also see Table S4 and Table S5). It is not immediately clear which epistatic subtypes are the most appropriate for these data, although including the minimum subtype may not be appropriate (Mani et al. 2008) (see discussion).Although it may be best to consider fewer epistatic subtypes for this specific data set, we report our results including all three epistatic subtypes and their corresponding null models (St. Onge et al. (2007), although we identify 105 epistatic pairs not identified by the original authors (Figure S4, Table S4). St. Onge et al. (2007) find that epistatic pairs with a functional link have a positively shifted distribution of epistasis. We find no such shift in epistasis values (Figure S5). We also demonstrate [described in application to simulated data: Bias and variance of the epistatic parameter (ɛ)] that our method seems to reduce bias of the epistatic parameter (ɛ) (Table S3).] When considering only a subset of the epistatic subtypes, however, we find to be positive and significantly different from zero (results not shown). See File S1, Figure S6, and Figure S7 for additional discussion of the epistatic pairs we identify.

Jasnos and Korona (2007) data set:

The Jasnos and Korona (2007) data set included 758 yeast gene deletions known to cause growth defects and reports fitnesses of only a sparse subset of all possible gene pairs [≈0.2% of the possible pairwise genotypes, or 639 pairs of ]. Because the authors do not identify epistatic pairs in a hypothesis-testing framework, we cannot explicitly compare our conclusions with theirs.To validate our method, we examine gene pairs that have specific functional links (see methods: Analysis of experimental data). When defining a functional link using GO terms (Gene Ontology Consortium 2000) with <30 genes associated with them, only 1 of 639 tested gene pairs has a functional link. Raising the threshold of associated genes to 50 and 100, the number of tested pairs with functional links rises only to 3 and 9, respectively. Because of the large number of random genes and the sparse number of gene pairs in this data set, we follow Tong et al. (2004) and select GO terms that have associated with them ≤200 genes. Twenty-five of 639 tested pairs then have a functional link.Only the FDR multiple-testing procedure results in a significant enrichment of functional links among epistatic pairs (File S1). With the FDR procedure we find 352 significant epistatic pairs, of which 35 (9.9%) have additive epistasis, 63 (17.9%) have multiplicative epistasis, and 254 (72.2%) have minimum epistasis (File S1_3). These proportions of inferred subtypes suggest that the authors'' original restriction to multiplicative epistasis may be inappropriate. We find 141 gene pairs with positive epistasis and 211 gene pairs with negative epistasis.We do not find a significant number of epistatic pairs with shared GO Slim Biological Process terms (see Analysis of experimental data), but do when considering all shared GO Slim terms (St. Onge et al. (2007) data set, we also consider some of the possible subsets of the three epistatic subtypes (and their corresponding null models) in our model selection procedure (Table S5). In contrast to the St. Onge et al. (2007) data set, using all three epistatic subtypes results in a significant number of epistatic pairs with functional links; this measure is not significant when using any of the other subsets of the subtypes. This suggests that our proposed method with three epistatic subtypes may indeed be the most appropriate for data sets with randomly selected genes.We examined the distribution of the estimated values of the epistatic parameter (ɛ) for all pairs with significant epistasis. Jasnos and Korona (2007), in assuming only multiplicative epistasis, conclude that epistasis is predominantly positive. However, we find that the estimated mean of epistasis is not significantly different from zero (two-sided t-test, P-value = 0.9578; Figure 1 and File S1.Open in a separate windowFigure 1.—Distribution of the epistasis values (ɛ) for significant epistatic pairs in the Jasnos and Korona (2007) data set, determined using the FDR procedure and the BIC method including all three epistatic subtypes and their corresponding null models. Mean of ɛ is −0.0009, with a standard deviation of 0.3177; median value is −0.0587. A similar plot is shown in Figure 3 of Jasnos and Korona (2007).  相似文献   

15.
Imaging cell biology in live animals: Ready for prime time     
Roberto Weigert  Natalie Porat-Shliom  Panomwat Amornphimoltham 《The Journal of cell biology》2013,201(7):969-979
Time-lapse fluorescence microscopy is one of the main tools used to image subcellular structures in living cells. Yet for decades it has been applied primarily to in vitro model systems. Thanks to the most recent advancements in intravital microscopy, this approach has finally been extended to live rodents. This represents a major breakthrough that will provide unprecedented new opportunities to study mammalian cell biology in vivo and has already provided new insight in the fields of neurobiology, immunology, and cancer biology.The discovery of GFP combined with the ability to engineer its expression in living cells has revolutionized mammalian cell biology (Chalfie et al., 1994). Since its introduction, several light microscopy–based techniques have become invaluable tools to investigate intracellular events (Lippincott-Schwartz, 2011). Among them are: time-lapse confocal microscopy, which has been instrumental in studying the dynamics of cellular and subcellular processes (Hirschberg et al., 1998; Jakobs, 2006; Cardarelli and Gratton, 2010); FRAP, which has enabled determining various biophysical properties of proteins in living cells (Berkovich et al., 2011); and fluorescence resonance energy transfer (FRET), which has been used to probe for protein–protein interactions and the local activation of specific signaling pathways (Balla, 2009). The continuous search for improvements in temporal and spatial resolution has led to the development of more sophisticated technologies, such as spinning disk microscopy, which allows the resolution of fast cellular events that occur on the order of milliseconds (Nakano, 2002); total internal reflection microscopy (TIRF), which enables imaging events in close proximity (100 nm) to the plasma membrane (Cocucci et al., 2012); and super-resolution microscopy (SIM, PALM, and STORM), which captures images with resolution higher than the diffraction limit of light (Lippincott-Schwartz, 2011).Most of these techniques have been primarily applied to in vitro model systems, such as cells grown on solid substrates or in 3D matrices, explanted embryos, and organ cultures. These systems, which are relatively easy to maintain and to manipulate either pharmacologically or genetically, have been instrumental in providing fundamental information about cellular events down to the molecular level. However, they often fail to reconstitute the complex architecture and physiology of multicellular tissues in vivo. Indeed, in a live organism, cells exhibit a 3D organization, interact with different cell types, and are constantly exposed to a multitude of signals originated from the vasculature, the central nervous system, and the extracellular environment. For this reason, scientists have been attracted by the possibility of imaging biological processes in live multicellular organisms (i.e., intravital microscopy [IVM]). The first attempt in this direction was in 1839, when Rudolph Wagner described the interaction of leukocytes with the walls of blood vessels in the webbed feet of a live frog by using bright-field transillumination (Wagner, 1839). Since then, this approach has been used for over a century to study vascular biology in thin areas of surgically exposed organs (Irwin and MacDonald, 1953; Zweifach, 1954) or by implanting optical windows in the skin or the ears (Clark and Clark, 1932). In addition, cell migration has also been investigated using transparent tissues, such as the fin of the teleost (Wood and Thorogood, 1984; Thorogood and Wood, 1987). The introduction of epifluorescence microscopy has enabled following in more detail the dynamics of individual cells in circulation (Nuttall, 1987), in tumors (MacDonald et al., 1992), or in the immune system (von Andrian, 1996), and the spatial resolution has been significantly improved by the use of confocal microscopy, which has made it possible to collect serial optical sections from a given specimen (Villringer et al., 1989; O’Rourke and Fraser, 1990; Jester et al., 1991). However, these techniques can resolve structures only within a few micrometers from the surface of optically opaque tissues (Masedunskas et al., 2012a). It was only in the early nineteen nineties, with the development of multiphoton microscopy, that deep tissue imaging has become possible (Denk et al., 1990; Zipfel et al., 2003b), significantly contributing to several fields, including neurobiology, immunology, and cancer biology (Fig. 1; Svoboda and Yasuda, 2006; Amornphimoltham et al., 2011; Beerling et al., 2011). In the last few years, the development of strategies to minimize the motion artifacts caused by the heartbeat and respiration has made it possible to successfully image subcellular structures with spatial and temporal resolutions comparable to those achieved in in vitro model systems, thus providing the opportunity to study cell biology in live mammalian tissues (Fig. 1; Weigert et al., 2010; Pittet and Weissleder, 2011).Open in a separate windowFigure 1.Spatial resolution and current applications of intravital microscopy. IVM provides the opportunity to image several biological processes in live animals at different levels of resolution. Low-magnification objectives (5–10×) enable visualizing tissues and their components under physiological conditions and measuring their response under pathological conditions. Particularly, the dynamics of the vasculature have been one of topic most extensively studied by IVM. Objectives with higher magnification (20–30×) have enabled imaging the behavior of individual cell over long periods of time. This has led to major breakthroughs in fields such as neurobiology, immunology, cancer biology, and stem cell research. Finally, the recent developments of strategies to minimize the motion artifacts caused by the heartbeat and respiration combined with high power lenses (60–100×) have opened the door to image subcellular structures and to study cell biology in live animals.The aim of this review is to highlight the power of IVM in addressing cell biological questions that cannot be otherwise answered in vitro, due to the intrinsic limitations of reductionist models, or by other more classical approaches. Furthermore, we discuss limitations and areas for improvement of this imaging technique, hoping to provide cell biologists with the basis to assess whether IVM is the appropriate choice to address their scientific questions.

Imaging techniques currently used to perform intravital microscopy

Confocal and two-photon microscopy are the most widely used techniques to perform IVM. Confocal microscopy, which is based on single photon excitation, is a well-established technique (Fig. 2 A) that has been extensively discussed elsewhere (Wilson, 2002); hence we will only briefly describe some of the main features of two-photon microscopy and other nonlinear optical techniques.Open in a separate windowFigure 2.Fluorescent light microscopy imaging techniques used for intravital microscopy. (A) Confocal microscopy. (top) In confocal microscopy, a fluorophore absorbs a single photon with a wavelength in the UV-visible range of the spectrum (blue arrow). After a vibrational relaxation (orange curved arrow), a photon with a wavelength shifted toward the red is emitted (green arrow). (center) In thick tissue, excitation and emission occur in a relative large volume around the focal plane (F.P.). The off-focus emissions are eliminated through a pinhole, and the signal from the focal plane is detected via a photomultiplier (PMT). Confocal microscopy enables imaging at a maximal depth to 80–100 µm. (bottom) Confocal z stack of the tongue of a mouse expressing the membrane marker m-GFP (green) in the K14-positive basal epithelial layer, and the membrane marker mTomato in the endothelium (red). The xy view shows a maximal projection of 40 z slices acquired every 2.5 µm, whereas the xz view shows a lateral view of the stack. In blue are the nuclei labeled by a systemic injection of Hoechst. Excitation wavelengths: 450 nm, 488 nm, and 562 nm. (B) Two- and three-photon microscopy. (top) In this process a fluorophore absorbs almost simultaneously two or three photons that have half (red arrow) or a third (dark red arrow) of the energy required for its excitation with a single photon. Two- or three-photon excitations typically require near-IR or IR light (from 690 to 1,600 nm). (center) Emission and excitation occur only at the focal plane in a restricted volume (1.5 fl), and for this reason a pinhole is not required. Two- and three-photon microscopy enable imaging routinely at a maximal depth of 300–500 µm. (bottom) Two-photon z stack of an area adjacent to that imaged in A. xy view shows a maximal projection of 70 slices acquired every 5 µm. xz view shows a lateral view of the stack. Excitation wavelength: 840 nm. (C) SHG and THG. (top) In SHG and THG, photons interact with the specimen and combine to form new photons that are emitted with twice or three times their initial energy without any energy loss. (center) These processes have similar features to those described for two- and three-photon microscopy and enable imaging at a maximal depth of 200–400 µm. (bottom) z stack of a rat heart excited by two-photon microscopy (740 nm) to reveal the parenchyma (green), and SHG (930 nm) to reveal collagen fibers (red). xy shows a maximal projection of 20 slices acquired every 5 µm. xz view shows a lateral view of the stack. Bars: (xy views) 40 µm; (xz views) 50 µm.The first two-photon microscope (Denk et al., 1990) was based on the principle of two-photon excitation postulated by Maria Göppert-Mayer in her PhD thesis (Göppert-Mayer, 1931). In this process a fluorophore is excited by the simultaneous absorption of two photons with wavelengths in the near-infrared (IR) or IR spectrum (from 690 to 1,600 nm; Fig. 2 B). Two-photon excitation requires high-intensity light that is provided by lasers generating very short pulses (in the femtosecond range) and is focused on the excitation spot by high numerical aperture lenses (Zipfel et al., 2003b). There are three main advantages in using two-photon excitation for IVM. First, IR light has a deeper tissue penetration than UV or visible light (Theer and Denk, 2006). Indeed, two-photon microscopy can resolve structures up to a depth of 300–500 µm in most of the tissues (Fig. 2 B), and up to 1.5 mm in the brain (Theer et al., 2003; Masedunskas et al., 2012a), whereas confocal microscopy is limited to 80–100 µm (Fig. 2 A). Second, the excitation is restricted to a very small volume (1.5 fl; Fig. 2 B). This implies that in two-photon microscopy there is no need to eliminate off-focus signals, and that under the appropriate conditions photobleaching and phototoxicity are negligible (Zipfel et al., 2003b). However, confocal microscopy induces out-of-focus photodamage, and thus is less suited for long-term imaging. Third, selected endogenous molecules can be excited, thus providing the contrast to visualize specific biological structures without the need for exogenous labeling (Zipfel et al., 2003a). Some of these molecules can also be excited by confocal microscopy using UV light, although with the risk of inducing photodamage.More recently, other nonlinear optical techniques have been used for IVM, and among them are three-photon excitation, and second and third harmonic generation (SHG and THG; Campagnola and Loew, 2003; Zipfel et al., 2003b; Oheim et al., 2006). Three-photon excitation follows the same principle as two-photon (Fig. 2 B), and can reveal endogenous molecules such as serotonin and melatonin (Zipfel et al., 2003a; Ritsma et al., 2013). In SHG and THG, photons interact with the specimen and combine to form new photons that are emitted with two or three times their initial energy (Fig. 2 C). SHG reveals collagen (Fig. 2 C) and myosin fibers (Campagnola and Loew, 2003), whereas THG reveals lipid droplets and myelin fibers (Débarre et al., 2006; Weigelin et al., 2012). Recently, two other techniques have been used for IVM: coherent anti-Stokes Raman scattering (CARS) and fluorescence lifetime imaging (FLIM). CARS that is based on two laser beams combined to match the energy gap between two vibrational levels of the molecule of interest, has been used to image lipids and myelin fibers (Müller and Zumbusch, 2007; Fu et al., 2008; Le et al., 2010). FLIM, which measures the lifetime that a molecule spends in the excited state, provides quantitative information on cellular parameters such as pH, oxygen levels, ion concentration, and the metabolic state of various biomolecules (Levitt et al., 2009; Provenzano et al., 2009; Bakker et al., 2012).We want to emphasize that two-photon microscopy and the other nonlinear techniques are the obligatory choice when the imaging area is located deep inside the tissue, endogenous molecules have to be imaged, or long-term imaging with frequent sampling is required. However, confocal microscopy is more suited to resolve structures in the micrometer range, because of the possibility of modulating the optical slice (Masedunskas et al., 2012a).

IVM to investigate biological processes at the tissue and the single cell level

The main strength of IVM is to provide information on the dynamics of biological processes that otherwise cannot be reconstituted in vitro or ex vivo. Indeed, IVM has been instrumental in studying several aspects of tissue physiopathology (Fig. 3, A and B). Although other approaches such as classical immunohistochemistry, electron microscopy, and indirect immunofluorescence may provide detailed structural and quantitative information on blood vessels, IVM enables measuring events such as variations of blood flow at the level of the capillaries or local changes in blood vessel permeability. These data have been instrumental in understanding the mechanisms of ischemic diseases and tumor progression, and in designing effective anticancer treatments.

Table 1.

IVM to study tissue physiopathology
EventOrganProbesReference
Measurements of local blood flow and glial cell functionBrainDextranHelmchen and Kleinfeld, 2008
Ischemia and reperfusionBrainSulphorhodamine 101, DextranZhang and Murphy, 2007; Masamoto et al., 2012;
Glomerular filtration and tubular reabsorptionKidneyDextran, AlbuminKang et al., 2006; Yu et al., 2007; Camirand et al., 2011
Blood flow patternsPancreatic isletsDextranNyman et al., 2008
Capillary response and synaptic activationOlfactory bulbDextranChaigneau et al., 2003
Imaging angiogenesis during wound healingSkullcapDextranHolstein et al., 2011
Pulmonary microvasculature and endothelial activationLungDextranPresson et al., 2011
Morphology of blood vessels and permeability in tumorsXenograftsDextran, RGD quantum dotsTozer et al., 2005; Smith et al., 2008; Vakoc et al., 2009; Fukumura et al., 2010
Hepatic transport into the bile canaliculiLiverCarboxyfluorescein diacetate Rhodamine 123Babbey et al., 2012; Liu et al., 2012
Progression of amyloid plaques in Alzheimer’s diseaseBrainCurcumin and metoxy-04Spires et al., 2005; Garcia-Alloza et al., 2007
Mitochondrial membrane potentialLiverTetramethylrhodamine methyl ester Rhodamine 123Theruvath et al., 2008; Zhong et al., 2008
Oxygen consumptionLiverRu(phen3)2+Paxian et al., 2004
Sarcomere contraction in humansSkeletal muscleEndogenous fluorescenceLlewellyn et al., 2008
Open in a separate windowOpen in a separate windowFigure 3.Imaging tissues and individual cells in live animals. (A) The vasculature of an immunocompromised mouse was highlighted by the systemic injection of 2 MD dextran (red) before (left) and after (right) the implant of breast cancer cells in the back (green). Note the change in shape of the blood vessels and their increased permeability (arrow). Images were acquired by two-photon microscopy (excitation wavelength: 930 nm). (B) The microvasculature in the liver of a mouse expressing the membrane marker mTomato (red) was highlighted by the injection of cascade blue dextran (blue) and imaged by confocal microscopy (excitation wavelengths: 405 nm and 561 nm). Note the red blood cells that do not uptake the dye and appear as dark objects in the blood stream (arrow). (C) Metastatic and nonmetastatic human adenocarcinoma cells were injected in the tongue of an immunocompromised mouse and imaged for four consecutive days by using two-photon microscopy (excitation wavelength: 930 nm). The metastatic cells, which express the fluorescent protein mCherry (red), migrate away from the edge of the tumor (arrows), whereas the nonmetastatic cells, which express the fluorescent protein Venus (green), do not. (D) A granulocyte moving inside a blood vessel in the mammary gland of a mouse expressing GFP-tagged myosin IIb (green) and labeled with the mitochondrial vital dye MitoTracker (red) was imaged in time lapse by using confocal microscopy (excitation wavelengths: 488 nm and 561 nm). Figure corresponds to Video 1. Time is expressed as minutes:seconds. Bars: (A) 100 µm; (B) 10 µm; (C) 30 µm; (D) 10 µm.IVM has also been used successfully to study the dynamics and the morphological changes of individual cells within a tissue (EventOrganProbeReferenceNeuronal morphology of hippocampal neuronsBrainThy1-GFP mice, dextranBarretto et al., 2011Neuronal circuitryBrainBrainbow miceLivet et al., 2007Dendritic spine development in the cortexBrainYFP H-line micePan and Gan, 2008Calcium imaging in the brainBrainGCAMPZariwala et al., 2012Natural killer cell and cytotoxic T cell interactions with tumorsXenograftmCFP , mYFPDeguine et al., 2010Neutrophil recruitment in beating heartHeartDextran, CX3CR1-GFP miceLi et al., 2012Immune cells in the central nervous systemBrainDextran, CX3CR1-GFP, LysM-GFP and CD11c-YFP miceNayak et al., 2012Dendritic cells migrationSkinYFP, VE-caherin RFP mice, dextranNitschké et al., 2012CD8+ T cells interaction with dendritic cells during viral infectionLymph nodesEGFP, Dextran, SHGHickman et al., 2008B cells and dendritic cells interactions outside lymph nodesLymph nodesEGFPQi et al., 2006Change in gene expression during metastasisXenograftPinner et al., 2009Invasion and metastasis in head and neck cancerXenograftYFP, RFP-lifeact, dextranAmornphimoltham et al., 2013Fibrosarcoma cell migration along collagen fibersDorsal skin chamberSHG, EGFP, DsRed, DextranAlexander et al., 2008Long term imaging mammary tumors and photo-switchable probesMammary windowDendra-2Kedrin et al., 2008; Gligorijevic et al., 2009Long term imaging liver metastasis through abdominal windowLiverSHG, Dendra2, EGFPRitsma et al., 2012bMacrophages during intravasation in mammary tumorsXenograftEGFP, SHG, dextransWang et al., 2007; Wyckoff et al., 2007Melanoma collective migrationDorsal skin ChamberSHG, THG, EGFP, DextranWeigelin et al., 2012Hematopoietic stem cells and blood vesselSkullcupDextranLo Celso et al., 2009Epithelial stem cells during hair regenerationSkinH2B-GFP miceRompolas et al., 2012Open in a separate windowIn neurobiology, for example, the development of approaches to perform long-term in vivo imaging has permitted the correlation of changes in neuronal morphology and neuronal circuitry to pathological conditions such as stroke (Zhang and Murphy, 2007), tumors (Barretto et al., 2011), neurodegenerative diseases (Merlini et al., 2012), and infections (McGavern and Kang, 2011). This has been accomplished by the establishment of surgical procedures to expose the brain cortex, and the implantation of chronic ports of observations such as cranial windows and imaging guide tubes for micro-optical probes (Svoboda and Yasuda, 2006; Xu et al., 2007; Barretto et al., 2011). In addition, this field has thrived thanks to the development of several transgenic mouse models harboring specific neuronal populations expressing either one or multiple fluorescent molecules (Svoboda and Yasuda, 2006; Livet et al., 2007).In tumor biology, the ability to visualize the motility of cancer cells within a tumor in vivo has provided tremendous information on the mechanisms regulating invasion and metastasis (Fig. 3 C; Beerling et al., 2011). Tumor cells metastasize to distal sites by using a combination of processes, which include tumor outgrowth, vascular intravasation, lymphatic invasion, or migration along components of the extracellular matrix and nerve fibers. Although classical histological analysis and indirect immunofluorescence have been routinely used to study these processes, the ability to perform long-term IVM through the optimization of optical windows (Alexander et al., 2008; Kedrin et al., 2008; Gligorijevic et al., 2009; Ritsma et al., 2012b) has provided unique insights. For example, a longitudinal study performed by using a combination of two-photon microscopy, SHG, and THG has highlighted the fact that various tissue components associated with melanomas may play either a migration-enhancing or migration-impeding role during collective cell invasion (Weigelin et al., 2012). In mammary tumors, the intravasation of metastatic cells has been shown to require macrophages (Wang et al., 2007; Wyckoff et al., 2007). In head and neck cancer, cells have been shown to migrate from specific sites at the edge of the tumor, and to colonize the cervical lymph nodes by migrating though the lymphatic vessels (Fig. 3 C; Amornphimoltham et al., 2013). In highly invasive melanomas, the migratory ability of cells has been correlated with their differentiation state, as determined by the expression of a reporter for melanin expression (Pinner et al., 2009).Imaging the cells of the immune system in a live animal has revealed novel qualitative and quantitative aspects of the dynamics of cellular immunity (Fig. 2 C and Video 1; Germain et al., 2005; Cahalan and Parker, 2008; Nitschke et al., 2008). Indeed, the very complex nature of the immune response, the involvement of a multitude of tissue components, and its tight spatial and temporal coordination clearly indicate that IVM is the most suited approach to study cellular immunity. This is highlighted in studies either in lymphoid tissues, where the exquisite coordination between cell–cell interactions and cell signaling has been studied during the interactions of B lymphocytes and T cell lymphoid tissues (Qi et al., 2006), T cell activation (Hickman et al., 2008; Friedman et al., 2010), and migration of dendritic cells (Nitschké et al., 2012), or outside lymphoid tissues, such as, for example, brain during pathogen infections (Nayak et al., 2012), heart during inflammation (Li et al., 2012), and solid tumors (Deguine et al., 2010).

Imaging subcellular structures in vivo and its application to cell biology

The examples described so far convey that IVM has contributed to unraveling how the unique properties of the tissue environment in vivo significantly regulate the dynamics of individual cells and ultimately tissue physiology. Is IVM suitable to determine (1) how subcellular events occur in vivo, (2) whether they differ in in vitro settings, and (3), finally, the nature of their contribution to tissue physiology?IVM has been extensively used to image subcellular structures in smaller organisms (i.e., zebrafish, Caenorhabditis elegans) that are transparent and can be easily immobilized (Rohde and Yanik, 2011; Tserevelakis et al., 2011; Hove and Craig, 2012). In addition, the ability to easily perform genetic manipulations has made these systems extremely attractive to study several aspects of developmental and cell biology. However, their differences in term of organ physiology with respect to rodents do not make them suitable models for human diseases. For a long time, subcellular imaging in live rodents has been hampered by the motion artifacts derived from the heartbeat and respiration. Indeed, small shifts along the three axes make it practically impossible to visualize structures whose sizes are in the micrometer or submicrometer range, whereas it marginally affects larger structures. This issue has been only recently addressed by using a combination of strategies, which include: (1) the development of specific surgical procedures that allow the exposure and proper positioning of the organ of interest (Masedunskas et al., 2013), (2) the improvement of specific organ holders (Cao et al., 2012; Masedunskas et al., 2012a), and (3) the synchronization of the imaging acquisition with the heartbeat and respiration (Presson et al., 2011; Li et al., 2012). Very importantly, these approaches have been successfully implemented without compromising the integrity and the physiology of the tissue, thus opening the door to study cell biology in a live animal.For example, large subcellular structures such as the nuclei have been easily imaged, making it possible to study processes such as cell division and apoptosis (Fig. 4 A; Goetz et al., 2011; Orth et al., 2011; Rompolas et al., 2012). Interestingly, these studies have highlighted the fact that the in vivo microenvironment substantially affects nuclear dynamics. Indeed, mitosis and the structure of the mitotic spindle were followed over time in a xenograft model of human cancer expressing the histone marker mCherry-H2B and GFP-tubulin (Orth et al., 2011). Specifically, the effects of the anticancer drug Paclitaxel were studied, revealing that the tumor cells in vivo have a higher mitotic index and lower pro-apoptotic propensity than in vitro (Orth et al., 2011). FRET has been used in subcutaneous tumors to image cytotoxic T lymphocyte–induced apoptosis and highlighted that the kinetics of this process are much slower than those reported for nontumor cells in vivo that are exposed to a different microenvironment (Breart et al., 2008). Cell division has also been followed in the hair-follicle stem cells of transgenic mice expressing GFP-H2B. This study determined that epithelial–mesenchymal interactions are essential for stem cell activation and regeneration, and that nuclear divisions occur in a specific area of the hair follicles and are oriented toward the axis of growth (Rompolas et al., 2012). These processes show an extremely high level of temporal and spatial organization that can only be appreciated in vivo and by using time-lapse imaging.Open in a separate windowFigure 4.Imaging subcellular events in live animals. (A) Human squamous carcinoma cells were engineered to stably express the Fucci cell cycle reporter into the nucleus and injected in the back of an immunocompromised mouse. After 1 wk, the tumor was imaged by two-photon microscopy and SHG (excitation wavelength: 930 nm). (top) Maximal projection of a z stack (xy view). Cells in G2/M are in green, cells in G1 are in red, and collagen fibers are in cyan. (bottom) Lateral view (xz) of a z stack. (B) Clusters of GLUT4-containing vesicles (green) in the soleus muscle of a transgenic mouse expressing GFP-GLUT4 and injected with 70 kD Texas red–dextran to visualize the vasculature and imaged by two-photon microscopy (excitation wavelength: 930 nm). (C) Confocal microscopy (excitation wavelength: 488 nm) of hepatocytes in the liver of a transgenic mouse expressing the autophagy marker GFP-LC3. The inset shows small GFP-LC3 autophagic vesicles. (D–G) Dynamics of intracellular compartments imaged by time-lapse two-photon (E) or confocal microscopy (D, F, and G). (D) Endocytosis of systemically injected 10 kD Texas red–dextran into the kidney of a transgenic mouse expressing the membrane marker m-GFP. The dextran (red) is transported from the microvasculature into the proximal tubuli, and then internalized in small endocytic vesicles (arrows; Video 2). (E) Endocytosis of a systemically injected 10 kD of Alexa Fluor 488 dextran into the salivary glands of a live rat. The dextran (green) diffuses from the vasculature into the stroma, and it is internalized by stromal cells (insets). Collagen fibers (red) are highlighted by SHG. (F) Regulated exocytosis of large secretory granules in the salivary glands of a live transgenic mouse expressing cytoplasmic GFP. The GFP is excluded from the secretory granules and accumulates on their limiting membranes (arrows) after fusion with the plasma membrane (broken lines). The gradual collapse of an individual granule is highlighted in the insets. (G) Dynamics of mitochondria labeled with the membrane potential dye TMRM in the salivary glands of a live mouse. Time is expressed as minutes:seconds. Bars: (A) 40 µm; (B) 15 µm; (C, D, E, and G) 10 µm; (F) 5 µm.Imaging membrane trafficking has been more challenging because of its dynamic nature and the size of the structures to image. The first successful attempt to visualize membrane traffic events was achieved in the kidney of live rats by using two-photon microscopy where the endocytosis of fluid-phase markers, such as dextrans, or the receptor-mediated uptake of folate, albumin, and the aminoglycoside gentamicin were followed in the proximal tubuli (Fig. 4 D and Video 2; Dunn et al., 2002; Sandoval et al., 2004; Russo et al., 2007). These pioneering studies showed for the first time that apical uptake is involved in the filtration of large molecules in the kidney, whereas previously it was believed to be exclusively due to a barrier in the glomerular capillary wall. However, in the kidney the residual motion artifacts limited the imaging to short periods of time. Recently, the salivary glands have proven to be a suitable organ to study the dynamics of membrane trafficking by using either two-photon or confocal microscopy. Systemically injected dextrans, BSA, and transferrin were observed to rapidly internalize in the stromal cells surrounding the salivary gland epithelium in a process dependent on the actin cytoskeleton (Masedunskas and Weigert, 2008; Masedunskas et al., 2012b). Moreover, the trafficking of these molecules through the endo-lysosomal system was documented, providing interesting insights on early endosomal fusion (Fig. 4 E; Masedunskas and Weigert, 2008; Masedunskas et al., 2012b). Notably, significant differences were observed in the kinetics of internalization of transferrin and dextran. In vivo, dextran was rapidly internalized by stromal cells, whereas transferrin appeared in endosomal structures after 10–15 min. However, in freshly explanted stromal cells adherent on glass, transferrin was internalized within 1 min, whereas dextran appeared in endosomal structures after 10–15 min. Although the reasons for this difference were not addressed, it is clear that the environment in vivo has profound effects on the regulation of intracellular processes (Masedunskas et al., 2012b). Similar differences have been reported for the caveolae that in vivo are more dynamic than in cell cultures (Thomsen et al., 2002; Oh et al., 2007). Endocytosis has also been investigated in the epithelium of the salivary glands (Sramkova et al., 2009). Specifically, plasmid DNA was shown to be internalized by a clathrin-independent pathway from the apical plasma membrane of acinar and ductal cells, and to subsequently escape from the endo-lysosomal system, thus providing useful information on the mechanisms of nonviral gene delivery in vivo (Sramkova et al., 2012). Receptor-mediated endocytosis has also been studied in cancer models. Indeed, the uptake of a fluorescent EGF conjugated to carbon nanotubes has been followed in xenografts of head and neck cancer cells revealing that the internalization occurs primarily in cells that express high levels of EGFR (Bhirde et al., 2009). The role of endosomal recycling has also been investigated during tumor progression. Indeed, the small GTPase Rab25 was found to regulate the ability of head neck cancer cells to migrate to lymph nodes by controlling the dynamic assembly of plasma membrane actin reach protrusion in vivo (Amornphimoltham et al., 2013). Interestingly, this activity of Rab25 was reconstituted in cells migrating through a 3D collagen matrix but not in cells grown adherent to a solid substrate.IVM has been a powerful tool in investigating the molecular machinery controlling regulated exocytosis in various organs. In salivary glands, the use of selected transgenic mice expressing either soluble GFP or a membrane-targeted peptide has permitted the characterization of the dynamics of exocytosis of the secretory granules after fusion with the plasma membrane (Fig. 4 E; Masedunskas et al., 2011a, 2012d). These studies revealed that the regulation and the modality of exocytosis differ between in vivo and in vitro systems. Indeed, in vivo, regulated exocytosis is controlled by stimulation of the β-adrenergic receptor, and secretory granules undergo a gradual collapse after fusion with the apical plasma membrane, whereas, in vitro, regulated exocytosis is also controlled by the muscarinic receptor and the secretory granules fuse to each other, forming strings of interconnected vesicles at the plasma membrane (compound exocytosis; Masedunskas et al., 2011a, 2012d). Moreover, the transient expression of reporter molecules for F-actin has revealed the requirement for the assembly of an actomyosin complex to facilitate the completion of the exocytic process (Masedunskas et al., 2011a, 2012d). This result underscores the fact that the dynamics of the assembly of the actin cytoskeleton can be studied both qualitatively and quantitatively in live animals at the level of individual secretory granules. In addition, this approach has highlighted some of the mechanisms that contribute to regulate the apical plasma membrane homeostasis in vivo that cannot be recapitulated in an in vitro model systems (Masedunskas et al., 2011b, 2012c; Porat-Shliom et al., 2013). Indeed, the hydrostatic pressure that is built inside the ductal system by the secretion of fluids that accompanies exocytosis plays a significant role in controlling the dynamics of secretory granules at the apical plasma membrane. This aspect has never been appreciated in organ explants where the integrity of the ductal system is compromised. Finally, a very promising model has been developed in the skeletal muscle, where the transient transfection of a GFP-tagged version of the glucose transporter type 4 (GLUT4) has made possible to characterize the kinetics of the GLUT4-containing vesicles in resting conditions and their insulin-dependent translocation to the plasma membrane (Fig. 4 B; Lauritzen et al., 2008, 2010). This represents a very powerful experimental model that bridges together physiology and cell biology and has the potential to provide fundamental information on metabolic diseases.These examples underscore the merits of subcellular IVM to investigate specific areas of cell biology such as membrane trafficking, the cell cycle, apoptosis, and cytoskeletal organization. However, IVM is rapidly extending to other areas, such as cell signaling (Stockholm et al., 2005; Rudolf et al., 2006; Ritsma et al., 2012a), metabolism (Fig. 4 C; Débarre et al., 2006; Cao et al., 2012), mitochondrial dynamics (Fig. 4 F; Sun et al., 2005; Hall et al., 2013), or gene and protein expression (Pinner et al., 2009) that have just begun to be explored.

Future perspectives

IVM has become a powerful tool to study biological processes in live animals that is destined to have an enormous impact on cell biology. The examples described here give a clear picture of the broad applicability of this approach. In essence, we foresee that IVM is going to be the obligatory choice to study highly dynamic subcellular processes that cannot be reconstituted in vitro or ex vivo, or when a link between cellular events and tissue physiopathology is being pursued. In addition, IVM will provide the opportunity to complement and confirm data generated from in vitro studies. Importantly, the fact that in several instances confocal microscopy can be effectively used for subcellular IVM makes this approach immediately accessible to several investigators.In terms of future directions, we envision that other light microscopy techniques will soon become standard tools for in vivo studies, as shown by the recent application of FRET to study signaling (Stockholm et al., 2005; Rudolf et al., 2006; Breart et al., 2008; Ritsma et al., 2012a), and FRAP, which has been used in the live brain to measure the diffusion of α synuclein, thus opening the door to studying the biophysical properties of proteins in vivo (Unni et al., 2010). Moreover, super-resolution microscopy may be applied for imaging live animals, although this task may pose some challenges. Indeed, these techniques require: (1) the complete stability of the specimen, (2) extended periods of time for light collection, (3) substantial modifications to the existing microscopes, and (4) the generation of transgenic mice expressing photoactivatable probes.To reach its full potential, IVM has to further develop two main aspects: animal models and instrumentations. Indeed, a significant effort has to be invested in developing novel transgenic mouse models, which express fluorescently labeled reporter molecules. One example is the recently developed mouse that expresses fluorescently tagged lifeact. This model will provide the unique opportunity to study F-actin dynamics in vivo in the context of processes such as cell migration and membrane trafficking (Riedl et al., 2010). Moreover, the possibility of crossing these reporter mice with knockout animals will provide the means to further study cellular processes at a molecular level. Alternatively, reporter molecules or other transgenes that may perturb a specific cellular pathway can be transiently transfected into live animals in several ways. Indeed, the remarkable advancements in gene therapy have contributed to the development of several nonviral- and viral-mediated strategies for gene delivery to selected target organs. In this respect, the salivary glands and the skeletal muscle are two formidable model systems because either transgenes or siRNAs can be successfully delivered without any adverse reaction and expressed in a few hours. In terms of the current technical limitations of IVM, the main areas of improvement are the temporal resolution, the ability to access the organ of interest with minimal invasion, and the ability to perform long-term imaging. As for the temporal resolution, the issue has begun to be addressed by using two different approaches: (1) the use of spinning disk microscopy, as shown by its recent application to image platelet dynamics in live mice (Jenne et al., 2011); and (2) the development of confocal and two-photon microscopes equipped with resonant scanners that permit increasing the scanning speed to 30 frames per second (Kirkpatrick et al., 2012). As for accessing the organs, recently several microlenses (350 µm in diameter) have been inserted or permanently implanted into live animals, minimizing the exposure of the organs and the risk of affecting their physiology (Llewellyn et al., 2008). Finally, although some approaches for the long-term imaging of the brain, the mammary glands, and the liver have been developed, additional effort has to be devoted to establish chronic ports of observations in other organs.In conclusion, these are truly exciting times, and a new era full of novel discoveries is just around the corner. The ability to see processes inside the cells of a live animal is no longer a dream.

Online supplemental material

Video 1 shows time-lapse confocal microscopy of a granulocyte moving inside a blood vessel in the mammary gland of a mouse expressing GFP-tagged myosin IIb (green) and labeled with MitoTracker (red). Video 2 shows time-lapse confocal microscopy of the endocytosis of systemically injected 10 kD Texas red–dextran (red) into the kidney-proximal tubuli of a transgenic mouse expressing the membrane marker m-GFP (green). Online supplemental material is available at http://www.jcb.org/cgi/content/full/jcb.201212130/DC1.  相似文献   

16.
Immunomodulation by Mesenchymal Stem Cells in Veterinary Species     
Danielle D Carrade  Dori L Borjesson 《Comparative medicine》2013,63(3):207-217
Mesenchymal stem cells (MSC) are adult-derived multipotent stem cells that have been derived from almost every tissue. They are classically defined as spindle-shaped, plastic-adherent cells capable of adipogenic, chondrogenic, and osteogenic differentiation. This capacity for trilineage differentiation has been the foundation for research into the use of MSC to regenerate damaged tissues. Recent studies have shown that MSC interact with cells of the immune system and modulate their function. Although many of the details underlying the mechanisms by which MSC modulate the immune system have been defined for human and rodent (mouse and rat) MSC, much less is known about MSC from other veterinary species. This knowledge gap is particularly important because the clinical use of MSC in veterinary medicine is increasing and far exceeds the use of MSC in human medicine. It is crucial to determine how MSC modulate the immune system for each animal species as well as for MSC derived from any given tissue source. A comparative approach provides a unique translational opportunity to bring novel cell-based therapies to the veterinary market as well as enhance the utility of animal models for human disorders. The current review covers what is currently known about MSC and their immunomodulatory functions in veterinary species, excluding laboratory rodents.Abbreviations: AT, adipose tissue; BM, Bone marrow; CB, umbilical cord blood; CT, umbilical cord tissue; DC, dendritic cell; IDO, indoleamine 2;3-dioxygenase; MSC, mesenchymal stem cells; PGE2, prostaglandin E2; VEGF, vascular endothelial growth factorMesenchymal stem cells (MSC, alternatively known as mesenchymal stromal cells) were first reported in the literature in 1968.39 MSC are thought to be of pericyte origin (cells that line the vasculature)21,22 and typically are isolated from highly vascular tissues. In humans and mice, MSC have been isolated from fat, placental tissues (placenta, Wharton jelly, umbilical cord, umbilical cord blood), hair follicles, tendon, synovial membrane, periodontal ligament, and every major organ (brain, spleen, liver, kidney, lung, bone marrow, muscle, thymus, pancreas, skin).23,121 For most current clinical applications, MSC are isolated from adipose tissue (AT), bone marrow (BM), umbilical cord blood (CB), and umbilical cord tissue (CT; 11,87,99 Clinical trials in human medicine focus on the use of MSC both for their antiinflammatory properties (graft-versus-host disease, irritable bowel syndrome) and their ability to aid in tissue and bone regeneration in combination with growth factors and bone scaffolds (clinicaltrials.gov).131 For tissue regeneration, the abilities of MSC to differentiate and to secrete mediators and interact with cells of the immune system likely contribute to tissue healing (Figure 1). The current review will not address the specific use of MSC for orthopedic applications and tissue regeneration, although the topic is covered widely in current literature for both human and veterinary medicine.57,62,90

Table 1.

Tissues from which MSC have been isolated
Tissue source (reference no.)
SpeciesFatBone marrowCord bloodCord tissueOther
Cat1348356
Chicken63
Cow13812108
Dog973, 5978, 119139Periodontal ligament65
Goat66964
Horse26, 13037, 40, 12367130Periodontal ligament and gingiva88
Nonhuman primate28, 545
Pig1351147014, 20, 91
Rabbit1288032Fetal liver93
Sheep849542, 55
Open in a separate windowOpen in a separate windowFigure 1.The dual roles of MSC: differentiation and modulation of inflammation.Long-term studies in veterinary species have shown no adverse effects with the administration of MSC in a large number of animals.9,10,53 Smaller, controlled studies on veterinary species have shown few adverse effects, such as minor localized inflammation after MSC administration in vivo.7,15,17,45,86,92,98 Private companies, educational institutions, and private veterinary clinics (including Tufts University, Cummins School of Veterinary Medicine, University of California Davis School of Veterinary Medicine, VetStem, Celavet, Alamo Pintado Equine Medical Center, and Rood and Riddle Equine Hospital) offer MSC as a clinical treatment for veterinary species. Clinical uses include tendon and cartilage injuries, tendonitis, and osteoarthritis and, to a lesser extent, bone regeneration, spinal cord injuries, and liver disease in both large and small animals.38,41,113 Even with this broad clinical use, there have been no reports of severe adverse effects secondary to MSC administration in veterinary patients.  相似文献   

17.
A Problem With the Correlation Coefficient as a Measure of Gene Expression Divergence     
Vini Pereira  David Waxman  Adam Eyre-Walker 《Genetics》2009,183(4):1597-1600
The correlation coefficient is commonly used as a measure of the divergence of gene expression profiles between different species. Here we point out a potential problem with this statistic: if measurement error is large relative to the differences in expression, the correlation coefficient will tend to show high divergence for genes that have relatively uniform levels of expression across tissues or time points. We show that genes with a conserved uniform pattern of expression have significantly higher levels of expression divergence, when measured using the correlation coefficient, than other genes, in a data set from mouse, rat, and human. We also show that the Euclidean distance yields low estimates of expression divergence for genes with a conserved uniform pattern of expression.IT is now possible to measure the expression levels of thousands of genes in multiple tissues at multiple times. This has led to investigations into the evolution of gene expression and how the pattern of expression changes on a genomic scale. In some analyses, the evolution of expression is considered only within one tissue, but in many studies the evolution across multiple tissues is investigated. In this latter case, the evolution of an expression profile—a vector of expression levels of a gene across several tissues—is considered.Several different statistics have been proposed to measure the divergence between gene expression profiles. The two most popular measures are the Euclidean distance (Jordan et al. 2005; Kim et al. 2006; Yanai et al. 2006; Urrutia et al. 2008) and Pearson''s correlation coefficient (Makova and Li 2003; Huminiecki and Wolfe 2004; Yang et al. 2005; Kim et al. 2006; Liao and Zhang 2006a,b; Xing et al. 2007; Urrutia et al. 2008). The correlation coefficient is often subtracted from one, so that the statistic varies from zero, when there has been no expression divergence, to a maximum of two; we refer to this statistic as the Pearson distance. Here we describe a significant shortcoming of the Pearson distance that is not shared by the Euclidean distance.To investigate properties of these two measures of expression divergence, we compiled a data set of 2859 orthologous genes from human, mouse, and rat for which we had microarray expression data from nine homologous tissues: bone marrow, heart, kidney, large intestine, pituitary, skeletal muscle, small intestine, spleen, and thymus). The expression data for rat came from Walker et al. (2004), the mouse data from Su et al. (2004), and the human data from Ge et al. (2005). Each tissue experiment had two replicates in mouse, a varying number of replicates in rat, and one in humans; some genes were also matched by multiple probe sets. To obtain an average across experiments and probe sets we processed the data as follows:
  1. Raw CEL files of gene expression levels were obtained from the NCBI Gene Expression Omnibus database (http://www.ncbi.nlm.nih.gov/projects/geo/).
  2. The results from the mouse, rat, and human arrays were normalized separately using both the MAS5 (Affymetrix 2001) and the RMA algorithms (Irizarry et al. 2003) as implemented in Bioconductor (Gentleman et al. 2004). The results are qualitatively similar for the two normalization procedures, although recent analyses suggest that MAS5 normalization is generally better (Ploner et al. 2005; Lim et al. 2007).
  3. The expression of each gene within a tissue was averaged across experiments and probe sets.
We computed expression distances (ED) between orthologous gene expression profiles, for each of the three species comparisons, rat–mouse, rat–human, and mouse–human, according to the two different distance metrics, the Euclidean distance and the Pearson distance:(1)Here xij is the expression level of the gene under consideration in species i in tissue j, and is the average expression level of the gene in species i across tissues. Expression levels are known in a total of k tissues.Because expression levels are measured on different microarray platforms in the three species, we compute relative abundance (RA) values, before calculating the Euclidean distance (Liao and Zhang 2006a). The RA is the expression of a gene in a particular tissue divided by the sum of the expression values of that gene across all tissues. We calculated RA values to remove “probe” effects (the tendency for a gene to bind its probe set on one platform more efficiently than on another platform). Because of probe effects it is not easy to distinguish absolute changes in expression and differences in binding efficiency. Calculating RA values removes this problem from the Euclidean distance. Pearson''s distance does not change under such a rescaling and so this is unnecessary.In some analyses the logarithm of the expression or RA values are used (e.g., Makova and Li 2003; Kim et al. 2006; Xing et al. 2007), and in others the expression values are used without this transformation (e.g., Huminiecki and Wolfe 2004; Jordan et al. 2005; Yang et al. 2005; Liao and Zhang 2006a,b; Yanai et al. 2006; Urrutia et al. 2008). We calculated both the Pearson and the Euclidean distances on log-transformed and untransformed expression values. The results are qualitatively similar so here we present only the results obtained using the logarithm of the expression or RA values.It is natural to expect the two measures of expression divergence to be positively correlated with one another; however, the Euclidean and Pearson distances are almost completely uncorrelated (MAS5 normalization, mouse–rat correlation coefficient = 0.06, human–rat r = 0.13, human–mouse r = 0.10; RMA normalization, mouse–rat correlation coefficient = −0.12, human–rat r = −0.00, human–mouse r = −0.08; Figure 1). This could, plausibly, be because the two statistics measure different aspects of divergence. However, irrespective of this, there is a potential problem associated with the Pearson distance. Imagine that we have a gene that is expressed at identical levels in all tissues in two species (i.e., expression levels are uniform between tissues and also between species). We quite reasonably assume that measured expression levels contain noise. Thus each measured expression level (xij) is the sum of the (assumed) uniform expression level and an independent random number representing noise. In this case there is no real divergence in the expression profile between the species. However, the two measures of divergence may differ greatly in this case. The Euclidean distance reflects only the noise present in the data and hence will be small if the noise is small. By contrast, the Pearson distance will have a value close to 1 since the second term in PeaD in Equation 1 will be close to zero, reflecting the fact that the noise components of different expression levels are independent. Thus the Pearson distance will give the impression that expression divergence is great, but all this apparent divergence is noise. This will be a problem with Pearson''s distance whenever measurement error is of the same magnitude as the differences in expression between tissues. This will therefore tend to be a problem for lowly expressed genes, where measurement error can be large relative to the true value.Open in a separate windowFigure 1.—The correlation between the Euclidean and Pearson distances for (a) mouse–rat, (b) human–rat, and (c) human–mouse. Only the results from MAS5 normalization are shown; qualitatively similar results were obtained with RMA.The above example is unrealistic because real gene expression profiles are rarely perfectly uniform. To investigate whether this shortcoming of the Pearson distance is a problem in real data sets, we determined genes with a relatively uniform pattern of expression in all three species considered above. To do this we computed the entropy of a gene''s expression, which is a measure of uniformity in expression across tissues (Schug et al. 2005): the higher the value of the entropy, the more uniform is the expression. We calculated the entropy for each gene in each of the three species, averaged these across species, and then took those genes in the upper quartile of mean entropy values as a data set of genes with a relatively conserved pattern of uniform expression.It is natural to expect those genes with a conserved uniform pattern of expression to have relatively low expression divergence; however, on average these genes have significantly higher Pearson distances than other genes (Figure 2; supporting information, Figure S1 and Figure S2). By contrast, the Euclidean distance shows the pattern one would anticipate; all of the conserved uniform genes have low expression divergence. It therefore seems likely that the Pearson distance is sensitive to measurement error and hence may not be a good measure of expression divergence.Open in a separate windowFigure 2.—The distribution of expression divergence values for those genes with a uniform pattern of expression that is conserved across species vs. the distribution for all genes for (a) Pearson and (b) Euclidean distances for mouse–rat. We present similar values for human–mouse and human–rat in Figure S1 and Figure S2. Only the results from MAS5 normalization are shown; qualitatively similar results were obtained with RMA.

TABLE 1

The median expression divergence for genes that have a conserved uniform pattern of expression (upper quartile of mean entropy values) vs. all other genes
Data setStatisticConserved uniform genesOther genesWilcoxon test P-value
MAS5 normalization
    Mouse–ratEuclidean1.662.79<10−15
Pearson0.700.47<10−15
    Human–mouseEuclidean1.673.13<10−15
Pearson0.780.58<10−15
    Human–ratEuclidean1.833.21<10−15
Pearson0.780.58<10−15
RMA normalization
    Mouse–ratEuclidean0.591.40<10−15
Pearson0.820.38<10−15
    Human–mouseEuclidean0.591.58<10−15
Pearson0.810.48<10−15
    Human–ratEuclidean0.581.55<10−15

Pearson
0.73
0.50
<10−15
Open in a separate windowWe note that there are two additional advantages of the Euclidean distance. First, it can take into account differences in the absolute level of expression if those data are available, either because the method of assay allows this, for example, if ESTs, SAGE, sequencing, or RNA-Seq data are used, or because expression in the two species has been assessed on the same platform using probes that are conserved between the two species. Second, the square of the Euclidean distance is expected to increase linearly with time. Khaitovich et al. (2004) have previously shown that the squared difference in log expression level increases linearly with time under a Brownian motion model of gene expression evolution. It is therefore expected that the squared Euclidean distance will increase with time since the squared Euclidean distance is the sum of the squared differences across tissues. We prove this in File S1; we also show that this linearity holds, approximately, when relative abundance values are used (see also Pereira et al. 2009).  相似文献   

18.
Modeling the Hydraulics of Root Growth in Three Dimensions with Phloem Water Sources     
Brandy S. Wiegers  Angela Y. Cheer  Wendy K. Silk 《Plant physiology》2009,150(4):2092-2103
Primary growth is characterized by cell expansion facilitated by water uptake generating hydrostatic (turgor) pressure to inflate the cell, stretching the rigid cell walls. The multiple source theory of root growth hypothesizes that root growth involves transport of water both from the soil surrounding the growth zone and from the mature tissue higher in the root via phloem and protophloem. Here, protophloem water sources are used as boundary conditions in a classical, three-dimensional model of growth-sustaining water potentials in primary roots. The model predicts small radial gradients in water potential, with a significant longitudinal gradient. The results improve the agreement of theory with empirical studies for water potential in the primary growth zone of roots of maize (Zea mays). A sensitivity analysis quantifies the functional importance of apical phloem differentiation in permitting growth and reveals that the presence of phloem water sources makes the growth-sustaining water relations of the root relatively insensitive to changes in root radius and hydraulic conductivity. Adaptation to drought and other environmental stresses is predicted to involve more apical differentiation of phloem and/or higher phloem delivery rates to the growth zone.Plant growth involves water uptake by the cells and expansion of the cell walls under the resultant turgor (internal hydrostatic pressure). The water uptake and increase in cell volume are accompanied by nutrient and metabolite deposition. Thus, hydraulics of growth (i.e. the energies, conductivities, and fluxes of water in growing tissue) are fundamental to understanding primary plant growth. Quantitatively, the driving force for water movement in the plant, as in other porous media, is considered to be the gradient in water potential (Ψ), an energy per unit volume given in MPa. Thus, primary growth can be modeled by considering plant tissue to be a distributed sink for water, with low Ψ and/or high hydraulic conductivity driving water deposition into rapidly expanding regions. Molz and Boyer (1978) developed the theoretical basis for predicting the radial water flux in one dimension within the intercalary meristem of growing soybean (Glycine max) hypocotyls. In this aerial tissue, water moves from the xylem both outward to the epidermis and inward to the pith. Thus, in the growing hypocotyls, Ψ is predicted to be least negative in the xylem and to decrease toward the epidermis and the pith. These predictions for growth-induced or growth-sustaining Ψ were confirmed when the experimental technology became sensitive enough to detect the gradients in Ψ (Nonami and Boyer, 1993). Passioura and Boyer (2003) expanded the theory to incorporate anatomical detail and corresponding spatial patterns of hydraulic conductivity. Their model explains experimental results on water relations during growth transients for many areas of the plant.The hydraulics of root growth differ from shoot growth because of differences in xylem anatomy. Root xylem becomes functional perhaps 1 cm behind the tip and well behind the growth zone. To enter the growing cells near the maize (Zea mays) root tip, externally supplied metabolites must move several millimeters without phloem (Fig. 1), and any water supplied by functional xylem would need to move more than 1 cm. Silk and Wagner (1980) provided a theoretical framework for a two-dimensional treatment of the growth-sustaining Ψ gradients in maize roots. They assumed that the water source was external (the soil or root-bathing medium) and that the root surface was in equilibrium with the soil or bathing medium, so that the flow path to growing cells in the root was predicted to be primarily inward. As in the shoot model, growing tissue was seen as a distributed sink for water. However, since the publication of that theory, experimental studies have revealed that the root tip is not in equilibrium with the bathing medium (Pritchard et al., 1996, 2000; Gould et al., 2004; Shimazaki et al., 2005). Pressure probes combined with osmotic potential determinations have shown that the Ψ of exterior root cells ranges from −0.17 to −0.6 MPa, depending on environmental conditions. This range is more negative than in the nutrient medium. Furthermore, evidence has accumulated that at least some water for root growth comes from the phloem. The most obvious evidence is perhaps the growth of nodal (adventitious) roots of maize, rice (Oryza sativa), and other gramineous plants (Westgate and Boyer, 1985). This growth is a normal part of crop development. The nodal roots grow through air and then dry layers of surface soil, making it unlikely that the expanding root cells obtain water from the dry media surrounding the root. Empirical and theoretical studies have concluded that the phloem probably provides water for growth of the primary maize root (Bret-Harte and Silk, 1994; Frensch and Hsiao, 1995; Pritchard, 1996; Pritchard et al., 1996, 2000; Hukin et al., 2002; Gould et al., 2004).Open in a separate windowFigure 1.Primary root growth zone. The tip of the seedling root of maize showing the meristem as part of the apical third of the elongation zone. The boundary of this root section was digitized to provide the computational body-fit grid used for the model. [See online article for color version of this figure.]The model described here follows the concepts of Pritchard and colleagues (1996, 2000) in assuming a pressure-driven bulk flow of solution through the phloem to the region where phloem is beginning to be functional (1–4 mm from the apex; Fig. 1). Water movement can occur from both the surrounding soil and the developing phloem. Henceforth, we refer to the “external water source equilibrium” or EE model, for which the boundary condition is solely an exterior medium of fairly high Ψ (−0.005 to −0.05 MPa) and no conditions are placed on the phloem Ψ (Silk and Wagner (1980), that the exterior of the root is in equilibrium with its bathing solution. Empirical studies have shown that this model is not realistic, because the root maintains peripheral cells at more negative Ψ than the bathing medium. Since this is hypothesized to occur by deposition of apoplastic solutes, we will refer to a model with external water source and apoplastic solutes near the exterior as the EASE model.

Table I.

Acronyms for models and definitions of symbols used in mathematical modeling
AcronymBoundary Condition
EEExternal water source Equilibrium
EASEExternal water source and Apoplastic Solutes near the Exterior
PEWSPhloem and External Water Sources
SymbolPhysical SignificanceUnits
LRelative elemental growth rate h−1
Growth velocity vectormm h−1
Water flux vectormm h−1
Hydraulic conductivity tensormm2 s−1 MPa−1
ΨTotal water potentialMPa
Unit normal to the surface
sControl surfacemm2
VControl volumemm3
rRadial coordinatemm
zLongitudinal coordinatemm
x, yCartesian coordinatesmm
JJacobian Matrix of Transformation
Open in a separate windowA “multiple source” model places boundary conditions on the Ψ of both the bathing medium and the phloem to simulate both external and internal source activity, so we will refer to this model as the PEWS (for phloem and external water sources) model.  相似文献   

19.
Enhancing the Activity of a Protein by Stereospecific Unfolding: CONFORMATIONAL LIFE CYCLE OF INSULIN AND ITS EVOLUTIONARY ORIGINS*S??     
Qing-xin Hua  Bin Xu  Kun Huang  Shi-Quan Hu  Satoe Nakagawa  Wenhua Jia  Shuhua Wang  Jonathan Whittaker  Panayotis G. Katsoyannis    Michael A. Weiss 《The Journal of biological chemistry》2009,284(21):14586-14596
A central tenet of molecular biology holds that the function of a protein is mediated by its structure. An inactive ground-state conformation may nonetheless be enjoined by the interplay of competing biological constraints. A model is provided by insulin, well characterized at atomic resolution by x-ray crystallography. Here, we demonstrate that the activity of the hormone is enhanced by stereospecific unfolding of a conserved structural element. A bifunctional β-strand mediates both self-assembly (within β-cell storage vesicles) and receptor binding (in the bloodstream). This strand is anchored by an invariant side chain (PheB24); its substitution by Ala leads to an unstable but native-like analog of low activity. Substitution by d-Ala is equally destabilizing, and yet the protein diastereomer exhibits enhanced activity with segmental unfolding of the β-strand. Corresponding photoactivable derivatives (containing l- or d-para-azido-Phe) cross-link to the insulin receptor with higher d-specific efficiency. Aberrant exposure of hydrophobic surfaces in the analogs is associated with accelerated fibrillation, a form of aggregation-coupled misfolding associated with cellular toxicity. Conservation of PheB24, enforced by its dual role in native self-assembly and induced fit, thus highlights the implicit role of misfolding as an evolutionary constraint. Whereas classical crystal structures of insulin depict its storage form, signaling requires engagement of a detachable arm at an extended receptor interface. Because this active conformation resembles an amyloidogenic intermediate, we envisage that induced fit and self-assembly represent complementary molecular adaptations to potential proteotoxicity. The cryptic threat of misfolding poses a universal constraint in the evolution of polypeptide sequences.How insulin binds to the insulin receptor (IR)2 is not well understood despite decades of investigation. The hormone is a globular protein containing two chains, A (21 residues) and B (30 residues) (Fig. 1A). In pancreatic β-cells, insulin is stored as Zn2+-stabilized hexamers (Fig. 1B), which form microcrystal-line arrays within specialized secretory granules (1). The hexamers dissociate upon secretion into the portal circulation, enabling the hormone to function as a zinc-free monomer. The monomer is proposed to undergo a change in conformation upon receptor binding (2). In this study, we investigated a site of conformational change in the B-chain (PheB24) (arrow in Fig. 1A). In classical crystal structures, this invariant aromatic side chain (tawny in Fig. 1B) anchors an antiparallel β-sheet at the dimer interface (blue in Fig. 1C). Total chemical synthesis is exploited to enable comparison of corresponding d- and l-amino acid substitutions at this site, an approach designated “chiral mutagenesis” (3-5). In the accompanying article, the consequences of this conformational change are investigated by photomapping of the receptor-binding surface (6). Together, these studies redefine the interrelation of structure and activity in a protein central to the hormonal control of metabolism.Open in a separate windowFIGURE 1.Sequence and structure of insulin. A, sequences of the B-chain (upper) and A-chain (lower) with disulfide bridges as indicated. The arrow indicates invariant PheB24. The B24-B28 β-strand is highlighted in blue. B, crystal structure of the T6 zinc insulin hexamer (Protein Data Bank code 4INS): ribbon model (left) and space-filling model (right). The B24-B28 β-strand is shown in blue, and the side chain of PheB24 is highlighted in tawny. The B-chain is otherwise dark gray; the A-chain, light gray; and zinc ions, magenta. Also shown at the left are the side chains of HisB10 at the axial zinc-binding sites. C, cylinder model of the insulin dimer showing the B24-B26 antiparallel β-sheet (blue) anchored by the B24 side chain (tawny circle). The A- and B-chains are shown in light and dark gray, respectively. The protomer at the left is shown in the R-state, in which the central α-helix of the B-chain is elongated (B3-B19 in the frayed Rf protomer of T3Rf3 hexamers and B1-B19 in the R protomer of R6 hexamers). The three types of zinc insulin hexamers share similar B24-B26 antiparallel β-sheets as conserved dimerization elements.The structure of an insulin monomer in solution resembles a crystallographic protomer (Fig. 2A) (7-9). The A-chain contains an N-terminal α-helix, non-canonical turn, and second helix; the B-chain contains an N-terminal segment, central α-helix, and C-terminal β-strand. The β-strand is maintained in an isolated monomer wherein the side chain of PheB24 (tawny in Fig. 2A), packing against the central α-helix of the B-chain, provides a “plug” to seal a crevice in the hydrophobic core (Fig. 2B). Anomalies encountered in previous studies of insulin analogs suggest that PheB24 functions as a conformational switch (4, 7, 10-14). Whereas l-amino acid substitutions at B24 generally impair activity (even by such similar residues as l-Tyr) (15), a seeming paradox is posed by the enhanced activities of nonstandard analogs containing d-amino acids (10-12).

TABLE 1

Previous studies of insulin analogs
AnalogAffinityaAssaybRef.
%
d-PheB24-insulin 180 Lymphocytes 10
l-AlaB24-insulin 1 Hepatocytes 68
l-AlaB24-insulin 3 Lymphocytes 69
d-PheB24-insulin 140 ± 9 Hepatocytes 11
l-AlaB24-insulin 1.0 ± 0.1 Hepatocytes 11
d-AlaB24-insulin 150 ± 9 Hepatocytes 11
GlyB24-insulin 78 ± 11 Hepatocytes 11
DKP-insulin 200c CHO cells 12
d-PheB24-DKP-insulin 180 CHO cells 12
l-AlaB24-DKP-insulin 7 CHO cells 12
GlyB24-DKP-insulin 50 CHO cells 12
Open in a separate windowaAffinities are given relative to wild-type insulin (100%).bLymphocytes are human, and hepatocytes are rat; CHO designates Chinese hamster ovary.cStandard deviations are not provided in this reference.Open in a separate windowFIGURE 2.Role of PheB24 in an insulin monomer. A, shown is a cylinder model of insulin as a T-state protomer. The C-terminal B-chain β-strand is shown in blue, and the PheB24 side chain is shown in tawny. The black portion of the N-terminal A-chain α-helix (labeled buried) indicates a hidden receptor-binding surface (IleA2 and ValA3). B, the schematic representation of insulin highlights the proposed role of the PheB24 side chain as a plug that inserts into a crevice at the edge of the hydrophobic core. C and D, whereas substitution of PheB24 by l-Ala (C) would only partially fill the B24-related crevice, its substitution by d-Ala (D) would be associated with a marked packing defect. An alternative conformation, designated the R-state, is observed in zinc insulin hexamers at high ionic strength (74) and upon binding of small cyclic alcohols (75) but has not been observed in an insulin monomer.Why do d-amino acid substitutions at B24 enhance the activity of insulin? In this study, we describe the structure and function of insulin analogs containing l-Ala or d-Ala at B24 (Fig. 2, C and D). Our studies were conducted within an engineered monomer (DKP-insulin, an insulin analog containing three substitutions in the B-chain: AspB10, LysB28, and ProB29) to circumvent effects of self-assembly (16). Whereas the inactive l-analog retains a native-like structure, the active d-analog exhibits segmental unfolding of the B-chain. Studies of corresponding analogs containing either l- or d-photoactivable probes (l-para-azido-PheB24 or d-para-azido-PheB24 (l- or d-PapB24), obtained from photostable para-amino-Phe (Pmp) precursors (17)) demonstrate specific cross-linking to the IR. Although photo-contacts map in each case to the N-terminal domain of the receptor α-subunit (the L1 β-helix), higher cross-linking efficiency is achieved by the d-probe. Together, this and the following study (6) provide evidence that insulin deploys a detachable arm that inserts between domains of the IR.Induced fit of insulin illuminates by its scope general principles at the intersection of protein structure and cell biology. Protein evolution is enjoined by multiple layers of biological selection. The pathway of insulin biosynthesis, for example, successively requires (a) specific disulfide pairing (in the endoplasmic reticulum), (b) subcellular targeting and prohormone processing (in the trans-Golgi network), (c) zinc-mediated protein assembly and microcrystallization (in secretory granules), and (d) exocytosis and rapid disassembly of insulin hexamers (in the portal circulation), in turn enabling binding of the monomeric hormone to target tissues (1). Each step imposes structural constraints, which may be at odds. This study demonstrates that stereospecific pre-detachment of a receptor-binding arm enhances biological activity but impairs disulfide pairing and renders the hormone susceptible to aggregation-coupled misfolding (18). Whereas the classical globular structure of insulin and its self-assembly prevent proteotoxicity (3, 19), partial unfolding enables receptor engagement. We envisage that a choreography of conformational change has evolved as an adaptative response to the universal threat of toxic protein misfolding.  相似文献   

20.
Evolutionary Strata in a Small Mating-Type-Specific Region of the Smut Fungus Microbotryum violaceum          下载免费PDF全文
Antonina A Votintseva  Dmitry A. Filatov 《Genetics》2009,182(4):1391-1396
DNA sequence analysis and genetic mapping of loci from mating-type-specific chromosomes of the smut fungus Microbotryum violaceum demonstrated that the nonrecombining mating-type-specific region in this species comprises ∼25% (∼1 Mb) of the chromosome length. Divergence between homologous mating-type-linked genes in this region varies between 0 and 8.6%, resembling the evolutionary strata of vertebrate and plant sex chromosomes.EVOLUTION of mating types or sex-determining systems often involves the suppression of recombination around the primary sex-determining or mating-type-determining locus. In animals and plants, it is often an entire or almost entire chromosome (Y or W in male or female heterogametic species, respectively) that ceases to recombine with its homologous (X or Z) chromosome (Charlesworth and Charlesworth 2000; Charlesworth 2008). Self-incompatibility loci in plants are also thought to be located in regions of suppressed recombination (Charlesworth et al. 2005; Kamau and Charlesworth 2005; Kamau et al. 2007; Li et al. 2007; Yang et al. 2007). Regardless of the phylogenetic position of a species, such nonrecombining regions are known to follow similar evolutionary trajectories. The nonrecombining region on the sex-specific chromosome expands in several steps, forming evolutionary strata—regions of different X/Y (or Z/W) divergence (Lahn and Page 1999; Handley et al. 2004; Sandstedt and Tucker 2004; Nicolas et al. 2005)—and genes in the nonrecombining regions gradually accumulate deleterious mutations that eventually render them dysfunctional (Charlesworth and Charlesworth 2005; Charlesworth 2008).Fungal mating-type systems are very diverse, with the number of mating types varying from two to several hundred (Casselton 2002). Like sex chromosomes in several animals and plants, suppressed recombination has evolved in regions near fungal mating-type loci, including in Ustilago hordei (Lee et al. 1999), Cryptococcus neoformans (Lengeler et al. 2002), and Neurospora tetrasperma (Menkis et al. 2008). These species have two mating types, but no morphologically distinct sexes. The mating-type locus (the region of suppressed recombination) of C. neoformans is small (∼100 kb) compared with known sex chromosomes and contains only ∼20 genes that, unlike many sex chromosomes (Y or W chromosomes), show no obvious signs of genetic degeneration (Lengeler et al. 2002; Fraser et al. 2004). Judging from the divergence between the homologous genes on the two mating-type-specific chromosomes, C. neoformans started to evolve sex chromosomes a long time ago because silent divergence between the two mating types in the most ancient region exceeds 100% (Fraser et al. 2004). Genes in the younger mating-type-specific region are much less diverged between the two sex chromosomes, suggesting that the evolution of the sex locus in C. neoformans might have proceeded through several steps. The nonrecombining region around the mating-type locus of N. tetrasperma is much larger than in C. neoformans (at least 6.6 Mb), and silent divergence between homologous genes on the mating-type-specific chromosomes ranges from zero to 9%, demonstrating that these mating-type-specific chromosomes evolved recently (Menkis et al. 2008).M. violaceum, which causes anther smut disease in Silene latifolia and other species in the family Caryophyllaceae, has two mating types, A1 and A2 (reviewed by Giraud et al. 2008), which are determined by the presence of mating-type-specific chromosomes (hereafter A1 and A2 chromosomes, or sex chromosomes) in the haploid stage of the life cycle (Hood 2002; Hood et al. 2004). The A1 and A2 chromosomes are distinguishable by size in pulsed-field electrophoresis, and it is possible to isolate individual chromosomes electrophoretically (Hood et al. 2004). Random fragments of A1 and A2 chromosomes have previously been isolated from mating-type-specific bands of pulsed-field separated chromosomes of M. violaceum (Hood et al. 2004). These fragments were assumed to be linked to mating type. The same method was used to isolate fragments of non-mating-type-specific chromosomes. On the basis of the analysis of their sequences, (Hood et al. 2004) proposed that mating-type-specific chromosomes in M. violaceum might be degenerate because they contained a lower proportion of protein-coding genes than other chromosomes. However, it was not determined whether the sequences isolated from the mating-type chromosomes originated from the mating-type-specific or from the recombining regions (Hood et al. 2004), and the relative sizes of these regions are not known for these M. violaceum chromosomes. We tested the mating-type specificity of 86 of these fragments and demonstrate that fewer than a quarter of these loci are located in the mating-type-specific region, suggesting that the nonrecombining region on the A1 and A2 chromosomes is quite small, while the rest of the chromosome probably recombines (like pseudoautosomal regions of sex chromosomes) and is therefore not expected to undergo genetic degeneration. Genetic mapping confirms the presence of two pseudoautosomal regions in the M. violaceum mating-type-specific chromosomes.As these chromosomes are mating type specific in the haploid stage of M. violaceum, mating-type-specific loci (or DNA fragments) can be identified by testing whether they are present exclusively in A1 or A2 haploid strains. We therefore prepared haploid A1 and A2 M. violaceum cultures from S. latifolia plants from two geographically remote locations (accessions Sl405 from Sweden and Sl127 from the French Pyrenees). Haploid sporidial cultures were isolated by a standard dilution method (Kaltz and Shykoff 1997; Oudemans and Alexander 1998). Mating types were determined by PCR amplification of each culture with primers designed for A1 and A2 pheromone receptor genes linked to A1 and A2 mating types (Yockteng et al. 2007). The primers were as follows: 5′-TGGCATCCCTCAATGTTTCC-3′ and 5′-CACCTTTTGATGAGAGGCCG-3′ for the A1 pheromone receptor (GenBank accession no. EF584742) and 5′-TGACGAGAGCATTCCTACCG-3′ and 5′-GAAGCGGAACTTGCCTTTCT-3′ for the A2 pheromone receptor (GenBank accession no. EF584741). Cultures with PCR product amplified only from an A1 or A2 pheromone receptor gene were selected for further use. The mating types of the cultures were verified by conjugating them in all combinations.The GenBank nucleotide database was searched using BLAST for sequences similar to those isolated by Hood et al. (2004). Sequences with similarity to transposable elements (TE) and other repeats were excluded. The resulting set of nonredundant sequences was used to design PCR primers for 98 fragments. Half of these were originally isolated from the A1 and half from the A2 chromosomes and are hereafter called A1-NNN or A2-NNN (where NNN is the locus number; supporting information, Table S1), which does not imply that these loci are A1 or A2 specific, but merely indicates that they were originally isolated from the A1 or A2 chromosomes. Amplification of these regions from new A1 and A2 M. violaceum cultures, independently isolated by ourselves, revealed that only 5 of the 49 loci isolated from the A1 chromosome are indeed A1 specific and only 6 of 49 isolated from the A2 chromosome are A2 specific. All other loci amplified from both A1 and A2 cultures. Figure 1 illustrates some of these results from the Swedish sample (Sl405).Open in a separate windowFigure 1.—Testing of mating-type specificity for loci isolated from A1 and A2 chromosomes. (a) PCR amplifications from haploid cultures from Sl405 using primers designed from six A1-originated loci. Loci in which a PCR product could be amplified only from A1 cultures (boxed) were classified as specific to mating type A1. (b) PCR tests of six A2-originated loci on the same set of haploids as in a. Loci in which a PCR product amplified only from A2 cultures (boxed) were classified as specific to mating type A2. Loci amplified from both A1 and A2 cultures are not mating type specific.The fragments that amplified from both A1 and A2 mating types may be in recombining regions, or they could be present in mating-type-specific regions on both A1 and A2 chromosomes. If they are in recombining regions, the A1- and A2-linked homologs should not be diverged from each other, but if they are in nonrecombining, mating-type-specific regions, the divergence of the A1- and A2-linked homologs should be roughly proportional to the time since recombination stopped in the region. We therefore sequenced and compared PCR fragments amplified from the two mating types of Sl405 or Sl127 cultures (GenBank accession nos. FI855822FI856001). Sequencing of PCR products showed that 12 (4 A1 and 8 A2) loci have more than one copy, and they were excluded from further analysis. Sequences of 61 loci were identical between the A1 and A2 strains, and four loci demonstrated low total divergence (0.24–0.61%) between the two mating types (otintseva and D. Filatov, unpublished results). Thus, these loci might be located in the recombining part of the mating-type-specific chromosomes. Ten of 75 loci that amplified in both mating types demonstrated multiple polymorphisms fixed between the mating types rather than between the locations. Given that the strains that we used in the analysis originated from two geographically distant locations, it is highly unlikely that multiple polymorphisms distinguishing the A1 and A2 sequences arose purely by chance; thus, these loci are probably located in the nonrecombining mating-type-specific region of the M. violaceum A1 and A2 chromosomes.

TABLE 1

Loci from mating-type-specific chromosomes of M. violaceum used for PCR analysis and genetic map construction
With nonzero A1/A2 divergenceb
LociMating type specific<1%>1%With zero A1/A2 divergencebTotal
A1a52 (1)3 (3)35 (3)45 (7)
A2a62 (0)7 (7)26 (3)41 (10)
Subtotal4 (1)10 (10)
Total1114 (11)61 (6)86 (17)
Open in a separate windowaA1, loci originated from the A1 sex chromosome; A2, loci originated from the A2 sex chromosome.bThe number of loci used for genetic map construction is in parentheses.To confirm the mating-type-specific or pseudoautosomal locations of the loci with and without A1/A2 divergence, we conducted genetic mapping in a family of 99 individuals, 50 of which were of mating type A1 and 49 of mating type A2. The family was generated by a cross between A1 and A2 M. violaceum strains from S. latifolia accessions Sl405 (Sweden) and Sl127 (France), respectively. The choice of strains from geographically distant locations was motivated by the hope of maximizing the number of DNA sequence differences between them that can be used as molecular genetic markers in segregation analysis. We inoculated S. latifolia seedlings with sporidial cultures of both mating types. For inoculation, petri dishes with 12-day-old seedlings of S. latifolia were flooded with 2.5 ml of inoculum suspension. Inoculum suspension consisted of equal volumes of the A1 and A2 sporidial cultures that were mixed and conjugated overnight at 14° under rotation (Biere and Honders 1996; Van Putten et al. 2003). Seedlings were potted 3 days after inoculation. Two months later, teliospores were collected from the flowers of the infected plant and grown in petri dishes on 3.6% potato dextrose agar medium. Haploid sporidia formed after meiosis were isolated and grown as separate cultures for DNA extraction. The mating types of single sporidia cultures were identified as described above. The loci analyzed in the segregation analysis were sequenced in the two parental haploid strains and in 99 (50 A1 and 49 A2) haploid strains that were generated in the cross. Single nucleotide differences between the parental strains were used as molecular genetic markers for segregation analysis in the progeny. The genetic map was constructed using MAPMAKER/EXP v3.0 (Lincoln et al. 1992) and MapDisto v1.7 (http://mapdisto.free.fr/).The resulting genetic map is shown in Figure 2. As expected, no recombination was observed between the 10 loci with diverged A1- and A2-linked copies. In addition, one marker with no A1/A2 divergence, A2-397, was also completely linked to the loci with significant A1/A2 divergence. This locus either may be very tightly linked to the nonrecombining mating-type-specific region or may have been added to that region more recently than the loci that had already accumulated some divergence between the alleles in the two mating types. The mating-type-specific pheromone receptor locus (Devier et al. 2009) and 11 mating-type-specific loci are also located in this nonrecombining region (Figure 2). Interestingly, the cluster of nonrecombining markers is flanked on both sides with markers that recombine in meiosis, demonstrating that there are pseudoautosomal regions on both ends of the mating-type-specific chromosomes.Open in a separate windowFigure 2.—Genetic map of the mating-type-determining chromosome in M. violaceum. Genetic distance (in centimorgans) and the relative positions of the markers are shown to the left and the right of the chromosome, respectively. The position of the nonrecombining region corresponds to the cluster of linked markers shown on the right of the figure. Total A1/A2 divergence is shown in parentheses. Eleven mating-type-specific markers (for which sequences are available from only one mating type), located in the nonrecombining mating-type-specific region, are not shown.Our results demonstrate that although the loci reported by Hood et al. (2004) were isolated from the A1 and A2 chromosomes, most of these loci are not located in the nonrecombining mating-type-specific regions. In fact, the nonrecombining region might be relatively small: of 86 tested fragments, only 21 appeared to be either mating type specific or linked to the mating-type locus. Assuming that these loci represent a random set of DNA fragments isolated from the A1 and A2 chromosomes, it is possible to estimate the size of the nonrecombining region using the binomial distribution: the nonrecombining region is expected to be 24.4% (95% CI: 16.7–33.6%) of the chromosome length. As the sizes of the A1 and A2 chromosomes are ∼3.4 and 4.2 Mb long (Hood 2002; Hood et al. 2004), the nonrecombining region might be ∼1 Mb long.Interestingly, total A1/A2 divergence for the 11 loci with A1- and A2-linked copies mapped to the nonrecombining region varied from 0% to 8.6% (Figure 2). In addition, 11 loci amplified from only one mating type. These genes could represent degenerated genes, some of which degenerated in A1 strains, and some in A2 strains. Alternatively, they might be highly diverged genes, such that the PCR primers amplify only one allele, and not the other. Variation in divergence may be the result of the stepwise cessation of recombination between the A1 and A2 chromosomes in M. violaceum, resembling the evolutionary strata reported for human, chicken, and white campion sex chromosomes (Lahn and Page 1999; Handley et al. 2004; Bergero et al. 2007). However, only the differences between the most and the least diverged loci are statistically significant (Devier et al. 2009), the M. violaceum mating-type region has at least three strata: one oldest stratum, including the pheromone receptor locus; a younger stratum with ∼5–9% A1/A2 divergence; and the youngest stratum with 1–4% divergence between the two mating types. There may also be an additional very recently evolved stratum containing the locus named A2-397, which is also present in all A1 strains tested, with no fixed differences between the A1 and A2 strains (No. of sites analyzedWithin A1
Within A2
Fixed differences between A1 and A2A1/A2 divergence (%)LociaSb totalSπ (%)cSπ (%)cA1/A2 divergence <1%A1-23645630020.4410.44A1-0456544000040.61A2-568413220.4820.4800.24A2-411480210.210010.31A1/A2 divergence >1%A1-2176679000091.35A1-12856990010.1881.49A1-199618130010.16122.02A2-4223449000092.62A2-516470140000142.98A2-404508200030.59173.64A2-4355062220.3920.39183.95A2-4734572310.2210.22214.81A2-4573031710.3300165.54A2-5755034750.9930.59398.55
Open in a separate windowaA1, loci originated from the A1 sex chromosome; A2, loci originated from the A2 sex chromosome.bS, number of polymorphic sites.cπ (%), average number of differences per 100 nucleotides.

TABLE 3

P-values for the 2 × 2 G-tests for significance of differences in A1/A2 divergence between the loci in the nonrecombining region
LaSbLocusA2-397A1-217A1-128A1-199A2-422A2-516A2-404A2-435A2-473A2-457
5190A2-397
6679A1-2170.006
5698A1-1280.0060.93
61812A1-1990.00070.410.48
3449A2-4220.00030.170.210.51
47014A2-5160.000030.060.0860.280.76
50817A2-40400.0250.0380.150.550.75
50618A2-43500.0150.0240.1040.450.620.86
45721A2-47300.0010.0030.01630.150.210.340.43
30316A2-45700.00090.00170.00970.090.130.2030.260.69
50339A2-5750000.000010.0020.0020.0030.0060.0550.199
Open in a separate windowP-values <0.05 are in boldface type.aL, the length of the region compared.bS, the number of nucleotide differences observed.As most of the loci isolated from the A1 and A2 chromosomes recombine in meiosis, they are not expected to degenerate. Thus, the observation of a higher proportion of TEs in these loci, compared to other chromosomes (Hood et al. 2004), is unlikely to reflect genetic degeneration attributable to a lack of recombination in these loci. A higher abundance of TEs in the sequences isolated from the A1 and A2 chromosomes, as reported by Hood et al. (2004), may simply reflect variation in the TE density across the genome. Thus, it remains to be seen whether M. violaceum mating-type-specific regions degenerate, similar to vertebrate Y (or W) chromosomes, or remain largely intact, as in C. neoformans (Lengeler et al. 2002). If the latter were the case, it may suggest that nonrecombining regions in fungi do not necessarily follow the same degenerative path as animal Y and W chromosomes. The analysis of sequences from the M. violaceum genome (and perhaps other fungal genomes) will hopefully provide the answer to this question.The lack of degeneration of mating-type-specific regions in C. neoformans may be due to the relatively small size of the nonrecombining regions. The 20 genes present in this region may not be sufficient for the operation of such detrimental population genetic processes as background selection or Muller''s ratchet because the speed of these processes depends critically on the number of active genes linked together (Charlesworth 2008). Larger mating-type-specific regions in M. violaceum might contain more genes; thus, more active genetic degeneration may be expected in this species. Indeed, many strains of M. violaceum show haplolethality linked to one of the mating types (Hood and Antonivics 2000; Thomas et al. 2003; Tellier et al. 2005), which may reflect the accumulation of deleterious mutations in the nonrecombining regions around the mating-type loci. Mating-type specificity of the markers that amplified in only A1 or A2 strains in this study may also reflect genetic degeneration.Another factor that may potentially prevent degeneration of genes linked to mating-type loci in fungi is the haploid expression of genes in these regions. In animals, many Y-linked genes have functional homologs on the X chromosome, and loss of the Y-linked gene may be compensated for by expression of the X-linked homologs. The haploid stage in an animal''s life cycle is very short, and very few genes are actively expressed in animal gametes (Schultz et al. 2003). In plants, on the other hand, a significant proportion of the genome is expressed in pollen (da Costa-Nunes and Grossniklaus 2003), and so the loss of Y-linked genes expressed in gametes may be more detrimental than in animals. Indeed, most genes isolated from the white campion X chromosome have intact Y-linked copies (Filatov 2005; Bergero et al. 2007), but due to the small number of genes available, it is still unclear whether genetic degeneration of Y-linked genes is indeed slower in this species (and in plants generally) compared to animal Y chromosomes. Haploid expression could be an even more powerful force in fungi and other organisms with haploid sexes, such as bryophytes, as most genes are expressed in the haploid stage. Further analysis of genetic degeneration in nonrecombining sex- or mating-type-specific regions in fungi and bryophytes will help to shed light on this question.  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